首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A semisynthetic peroxidase was designed by exploiting the structural similarity of the active sites of vanadium dependent haloperoxidases and acid phosphatases. Incorporation of vanadate ion into the active site of phytase (E.C. 3.1.3.8), which mediates in vivo the hydrolysis of phosphate esters, leads to the formation of a semisynthetic peroxidase, which catalyzes the enantioselective oxidation of prochiral sulfides with H(2)O(2) affording the S-sulfoxide, e.g. in 66% ee at 100% conversion for thioanisole. Under reaction conditions the semi-synthetic vanadium peroxidase is stable for over 3 days with only a slight decrease in turnover frequency. Polar water-miscible cosolvents, such as methanol, dioxane, and dimethoxyethane, can be used in concentrations of 30% (v/v) at a small penalty in activity and enantioselectivity. Among the transition metal oxoanions that are known to be potent inhibitors, only vanadate resulted in a semisynthetic peroxidase when incorporated into phytase. A number of other acid phosphatases and hydrolases were tested for peroxidase activity, when incorporated with vanadate ion. Phytases from Aspergillus ficuum, A. fumigatus, and A. nidulans, sulfatase from Helix pomatia, and phospholipase D from cabbage catalyzed enantioselective oxygen transfer reactions when incorporated with vanadium. However, phytase from A. ficuum was unique in also catalyzing the enantioselective sulfoxidation, albeit at a lower rate, in the absence of vanadate ion.  相似文献   

2.
The operational stability of peroxidases was considerably enhanced by generating hydrogen peroxide in situ from glucose and oxygen. For example, the total turnover number of microperoxidase-11 in the oxidation of thioanisole was increased sevenfold compared with that obtained with continuous addition of H(2)O(2). Coimmobilization of peroxidases with glucose oxidase into polyurethane foams afforded heterogeneous biocatalysts in which the hydrogen peroxide is formed inside the polymeric matrix from glucose and oxygen. The total turnover number of chloroperoxidase in the oxidation of thioanisole and cis-2-heptene was increased to new maxima of 250. 10(3) and 10. 10(3), respectively, upon coimmobilization with glucose oxidase. Soybean peroxidase, which normally shows only classical peroxidase activity, was transformed into an oxygen-transfer catalyst when coimmobilized with glucose oxidase. The combination catalyst mediated the enantioselective oxidation of thioanisole [50% ee (S)] with 210 catalyst turnovers.  相似文献   

3.
The vanadate cofactor in vanadium chloroperoxidase has been studied using UV-VIS absorption spectroscopy. A band is present in the near-UV that is red-shifted as compared to free vanadate and shifts in both position and intensity upon change in pH. Mutation of vanadate binding residues has a clear effect on the spectrum. Substrate-induced spectral effects allow direct measurement of separate kinetics steps for the first time for vanadium haloperoxidases. A peroxo intermediate is formed upon addition of H(2)O(2), which causes a decrease in the absorption spectrum at 315 nm, as well as an increase at 384 nm. This peroxo form is very stable at pH 8.3, whereas it is less stable at pH 5.0, which is the optimal pH for activity. Upon addition of halides to the peroxo form, the native spectrum is re-formed as a result of halide oxidation. Stopped-flow experiments show that H(2)O(2) binding and Cl(-) oxidation occur on the millisecond to second time scale. These data suggest that the oxidation of Cl(-) to HOCl occurs in at least two steps. In the presence of H(2)O(2), the affinity for the vanadate cofactor was found to be much higher than previously reported for vanadate in the absence of H(2)O(2). This is attributed to the uptake of pervanadate by the apo-enzyme. Human glucose-6-phosphatase, which is evolutionarily related to vanadium chloroperoxidase, is also likely to have a higher affinity for pervanadate than vanadate. This could explain the enhanced insulin mimetic effect of pervanadate as compared to vanadate.  相似文献   

4.
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   

5.
Vanadium haloperoxidases and the bacterial class A nonspecific acid phosphatases have a conserved active site. It is shown that vanadate-substituted recombinant acid phosphatase from Shigella flexneri (PhoN-Sf) and Salmonella enterica ser. typhimurium (PhoN-Se) in the presence of H2O2 are able to oxidize bromide to hypobromous acid. Vanadate is essential for this activity. The kinetic parameters for the artificial bromoperoxidases have been determined. The Km value for H2O2 is about the same as that for the vanadium bromoperoxidases from the seaweed Ascophyllum nodosum. However, the Km value for Br- is about 10-20 times higher, and the turnover values of about 3.4 min-1 and 33 min-1 for PhoN-Sf and PhoN-Se, respectively, are much slower, than those of the native bromoperoxidase. Thus, despite the striking similarity in the active-site structures of the vanadium haloperoxidases and the acid phosphatase, the turnover frequency is low, and clearly the active site of acid phosphatases is not optimized for haloperoxidase activity. Like the native vanadium bromoperoxidase, the vanadate-substituted PhoN-Sf and PhoN-Se catalyse the enantioselective sulfoxidation of thioanisole.  相似文献   

6.
We synthesized vanadyl (oxidation state +IV) and vanadate (oxidation state +V) complexes with the same hydroxamic acid derivative ligand, and assessed their glucose-lowering activities in relation to the vanadium biodistribution behavior in streptozotocin-induced diabetic mice. When the mice received an intraperitoneal injection of the complexes, the vanadate complex more effectively lowered the elevated glucose levels compared with the vanadyl one. The glucose-lowering effect of the vanadate complex was linearly related to its dose within the range from 2.5 to 7.5 mg V/kg. In addition, pretreatment of the vanadate complex induced a larger insulin-enhancing effect than the vanadyl complex. Both complexes were more effective than the corresponding inorganic vanadium compounds. The vanadyl and vanadate complexes, but not the inorganic vanadium compounds, resulted in almost the same organ vanadium distribution. Consequently, the observed differences in the insulin-like activity between the complexes would reflect the potency of the two compounds in the +IV and +V oxidation states in the subcellular region.  相似文献   

7.
In the elucidation of structural requirements of heme vicinity for hydrogen peroxide activation, we found that the replacement of His-64 of myoglobin (Mb) with a negatively charged aspartate residue enhanced peroxidase and peroxygenase activities by 78- and 580-fold, respectively. Since residue 68 is known to influence the ligation of small molecules to the heme iron, we constructed H64D/V68X Mb bearing Ala, Ser, Leu, Ile, and Phe at position 68 to improve the oxidation activity. The Val-68 to Leu mutation of H64D Mb accelerates the reaction with H(2)O(2) to form a catalytic species, called compound I, and improves the one-electron oxidation of 2,2'-azinobis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) (i.e., peroxidase activity) approximately 2-fold. On the other hand, H64D/V68I Mb oxygenates thioanisole 2.7- and 1600-fold faster than H64D and wild-type Mb, respectively. In terms of the enantioselectivity, H64D/V68A and H64D/V68S Mb were good chiral catalysts for thioanisole oxidation and produced the (R)-sulfoxide dominantly with 84% and 88% ee, respectively [Kato, S., et al. (2002) J. Am. Chem. Soc. 124, 8506-8507]. On the contrary, the substitution of Val-68 in H64D Mb with an isoleucine residue alters the dominant sulfoxide product from the (R)- to the (S)-isomer. The crystal structures of H64D/V68A and H64D/V68S Mb elucidated in this study do not clearly indicate residues interacting with thioanisole. However, comparison of the active site structures provides the basis to interpret the changes in oxidation activity: (1) direct steric interactions between residue 68 and substrates (i.e., H(2)O(2), ABTS, thioanisole) and (2) the polar interactions between tightly hydrogen-bonded water molecules and substrates.  相似文献   

8.
Recombinant cytochrome c peroxidase (CcP) and a W51A mutant of CcP, in contrast to other classical peroxidases, react with phenylhydrazine to give sigma-bonded phenyl-iron complexes. The conclusion that the heme iron is accessible to substrates is supported by the observation that CcP and W51A CcP oxidize thioanisole to the racemic sulfoxide with quantitative incorporation of oxygen from H2O2. Definitive evidence for an open active site is provided by stereoselective epoxidation by both enzymes of styrene, cis-beta-methylstyrene, and trans-beta-methylstyrene. trans-beta-methylstyrene yields exclusively the trans-epoxide, but styrene yields the epoxide and phenylacetaldehyde, and cis-beta-methylstyrene yields both the cis- and trans-epoxides and 1-phenyl-2-propanone. The sulfoxide, stereoretentive epoxides, and 1-phenyl-2-propanone are formed by ferryl oxygen transfer mechanisms because their oxygen atom derives from H2O2. In contrast, the oxygen in the trans-epoxide from the cis-olefin derives primarily from molecular oxygen and is probably introduced by a protein cooxidation mechanism. cis-[1,2-2H]-1-Phenyl-1-propene is oxidized to [1,1-2H]-1-phenyl-2-propanone without a detectable isotope effect on the epoxide:ketone product ratio. The phenyl-iron complex is not formed and substrate oxidation is not observed when the prosthetic group is replaced by delta-meso-ethylheme. CcP thus has a sufficiently open active site to form a phenyl-iron complex, to oxidize thioanisole to the sulfoxide, and to epoxidize styrene and beta-methylstyrene. The results indicate that a ferryl (Fe(IV) = O)/protein radical pair can be coupled to achieve two-electron oxidations. The unique ability of CcP to catalyze monooxygenation reactions does not conflict with its peroxidase function because cytochrome c is oxidized at a distinct surface site (DePillis, G. D., Sishta, B. P., Mauk, A. G., and Ortiz de Montellano, P. R. (1991) J. Biol. Chem. 266, 19334-19341).  相似文献   

9.
In order to estimate the size of the cavity remaining around the heme of the 3A3-microperoxidase 8 (MP8) hemoabzyme, the formation of 3A3-MP8-Fe(II)-nitrosoalkane complexes upon oxidation of N-monosubstituted hydroxylamines was examined. This constituted a new reaction for hemoabzymes and is the first example of fully characterized Fe(II)-metabolite complexes of antibody-porphyrin. Also, via a comparison of the reactions with N-substituted hydroxylamines of various size and hydrophobicity, antibody 3A3 was confirmed to bring about a partial steric hindrance on the distal face of MP8. Subsequently, the influence of the antibody on the stereoselectivity of the S-oxidation of sulfides was examined. Our results showed that MP8 alone and the antibody-MP8 complex catalyze the oxidation of thioanisole by H(2)O(2) and tert-butyl hydroperoxide, following a peroxidase-like two-step oxygen-transfer mechanism involving a radical-cation intermediate. The best system, associating H(2)O(2) as oxidant and 3A3-MP8 as a catalyst, in the presence of 5% tert-butyl alcohol, led to the stereoselective S-oxidation of thioanisole with a 45% enantiomeric excess in favour of the R isomer. This constitutes the highest enantiomeric excess reported to date for the oxidation of sulfides catalyzed by hemoabzymes.  相似文献   

10.
 The action of vanadate on intact human erythrocytes was studied by 1H spin echo and 51V NMR spectroscopy as a model for the behaviour of vanadium(V) complexes in experimental diabetes. Vanadate is reduced by the intact erythrocyte at the expense of intracellular glutathione which rapidly depletes from the intracellular volume. Using the blocking agent 4,4′-diisothio-cyanatostilbene-2,2′-disulfonic acid (DIDS), which specifically blocks the anion transporter, vanadate reduction could be inhibited and glutathione depletion arrested. Thus, for the reaction with the intact cell to occur, vanadium(V) must cross the cell wall, possibly via the anion transporter. Nitrofurantoin was used to inhibit glutathione reductase in the erythrocyte suspensions. Under these conditions, treatment of the cells with vanadate induced glutathione oxidation prior to depletion. A study of the reaction of vanadate with haemolysate indicates that, without the influence of the membrane, rapid oxidation of glutathione to glutathione disulfide by the vanadyl cation occurs with no glutathione depletion, and that under these conditions vanadate reduction is incomplete. This study generates a model for the behaviour of vanadium complexes in vivo, providing a basis for the rational design and synthesis of new vanadium-based agents as insulin mimics. In essence, vanadium is transported across the membrane as vanadate(V), is reduced in situ by glutathione, and becomes complexed to a wide range of intracellular binding sites. Exchange reactions between glutathione and sulfhydryl groups present on haemoglobin and membrane lead to the depletion of glutathione from the cytosol. Received: 12 June 1996 / Accepted: 20 January 1997  相似文献   

11.
Sugars and sugar phosphates enable vanadate to catalyze the oxidation of NADH. Superoxide dismutase inhibits this oxidation. Incubation of sugars with vanadate, prior to addition of NADH, accelerates this oxidation of subsequently added NADH and eliminates the lag phase otherwise noted. Incubation of sugars with vanadate also results in the reduction of vanadate to vanadyl, with appearance of a blue-green color probably associated with a vanadyl-vanadate complex. It appears that sugars reduce vanadate to vanadyl which, in turn, reduces O2 to O2- and that vanadate plus O2- then catalyzes the oxidation of NAD(P)H by a free radical chain reaction. Such oxidation of NAD(P)H may account for several of the biological effects of vanadate.  相似文献   

12.
Peroxidase-catalyzed asymmetric sulfoxidations, while synthetically attractive, suffer from relatively low reaction rates due to poor substrate solubilities in water and from appreciable spontaneous oxidation of substrates (especially aryl alkyl sulfides) with H(2)O(2). In this work, we found that both of these shortcomings could be alleviated by switching from aqueous solutions to certain nearly anhydrous (99.7%) organic solvents as sulfoxidation reaction media. The rates of spontaneous oxidation of the model prochiral substrate thioanisole in several organic solvents were observed to be some 100- to 1000-fold slower than in water. In addition, the rates of asymmetric sulfoxidation of thioanisole in isopropyl alcohol and in methanol catalyzed by horseradish peroxidase (HRP) were determined to be tens to hundreds of times faster than in water under otherwise identical conditions. This dramatic activation is due to a much higher substrate solubility in organic solvents than in water and occurs even though the intrinsic reactivity of HRP in isopropyl alcohol and in methanol is hundreds of times lower than in water. Sulfoxidation of thioanisole catalyzed by four other hemoproteins (soybean peroxidase, myoglobin, hemoglobin, and cytochrome c) is also much faster in isopropyl alcohol than in water.  相似文献   

13.
Among the biotargets interacting with vanadium is the calcium pump from the sarcoplasmic reticulum (SR). To this end, initial research efforts were launched with two vanadium(V)-citrate complexes, namely (NH(4))(6)[V(2)O(4)(C(6)H(4)O(7))(2)].6H(2)O and (NH(4))(6)[V(2)O(2)(O(2))(2)(C(6)H(4)O(7))(2)].4H(2)O, potentially capable of interacting with the SR calcium pump by combining kinetic studies with (51)V NMR spectroscopy. Upon dissolution in the reaction medium (concentration range: 4-0.5mM), both vanadium(V):citrate (VC) and peroxovanadium(V):citrate (PVC) complexes are partially converted into vanadate oligomers. A 1mM solution of the PVC complex, containing 184microM of the PVC complex, 94microM oxoperoxovanadium(V) (PV) species, 222microM monomeric (V1), 43microM dimeric (V2) and 53microM tetrameric (V4) species, inhibits Ca(2+) accumulation by 75 %, whereas a solution of the VC complex of the same vanadium concentration, containing 98microM of the VC complex, 263microM monomeric (V1), 64microM dimeric (V2) and 92microM tetrameric (V4) species inhibits the calcium pump activity by 33 %. In contrast, a 1 mM metavanadate solution, containing 460microM monomeric (V1), 90.2microM dimeric (V2) and 80microM tetrameric (V4) species, has no effect on Ca(2+) accumulation. The NMR signals from the VC complex (-548.0ppm), PVC complex (-551.5ppm) and PV (-611.1ppm) are broadened upon SR vesicle addition (2.5mg/ml total protein). The relative order for the half width line broadening of the NMR signals, which reflect the interaction with the protein, was found to be V4>PVC>VC>PV>V2=V1=1, with no effect observed for the V1 and V2 signals. Putting it all together the effects of two vanadium(V)-citrate complexes on the modulation of calcium accumulation and ATP hydrolysis by the SR calcium pump reflected the observed variable reactivity into the nature of key species forming upon dissolution of the title complexes in the reaction media.  相似文献   

14.
Vanadium K-edge X-ray absorption spectroscopy (XAS) has been used to track the uptake and fate of VO(2+) ion in blood cells from Ascidia ceratodes, following exposure to dithiothreitol (DTT) or to DTT plus VO(2+). The full range of endogenous vanadium was queried by fitting the XAS of blood cells with the XAS spectra of model vanadium complexes. In cells exposed only to DTT, approximately 0.4% of a new V(III) species was found in a site similar to Na[V(edta)(H(2)O)]. With exposure to DTT and VO(2+), average intracellular [VO(aq)](2+) increased from 3% to 5%, and 6% of a new complexed form of vanadyl ion appeared evidencing a ligand array similar to [VO(edta)](2-). At the same time, the relative ratio of blood cell [V(H(2)O)(6)](3+) increased at the expense of [V(H(2)O)(5)(SO(4))](+) in a manner consistent with a significant increase in endogenous acidity. In new UV/Visible experiments, VO(2+) could be reduced to 7-coordinate [V(nta)(H(2)O)(3)] or [V(nta)(ida)](2-) with cysteine methyl ester in pH 6.5 solution. Ascorbate reduced [VO(edta)](2-) to 7-coordinate [V(edta)(H(2)O)](-), while [VO(trdta)](2-) was unreactive. These results corroborate the finding that the reductive EMF of VO(2+) is increased by the availability of a 7-coordinate V(III) product. Finally, a new and complete hypothesis is proposed for an ascidian vanadate reductase. The structure of the enzyme active site, the vanadate-vanadyl-vanadic reduction mechanism, the cellular locale, and elements of the regulatory machinery governing the biological reduction of vanadate and vanadyl ion by ascidians are all predicted. Together these constitute the new field of vanadium redox enzymology.  相似文献   

15.
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   

16.
Witting PK  Mauk AG  Lay PA 《Biochemistry》2002,41(38):11495-11503
Myoglobin (Mb) catalyzes a range of oxidation reactions in the presence of hydrogen peroxide (H(2)O(2)) through a peroxidase-like cycle. C110A and Y103F variants of human Mb have been constructed to assess the effects of removing electron-rich oxidizable amino acids from the protein on the peroxidase activity of Mb: a point mutation at W14 failed to yield a viable protein. Point mutations at C110 and Y103 did not result in significant changes to structural elements of the heme pocket, as judged by low-temperature electron paramagnetic spectroscopy (EPR) studies on the ground-state ferric proteins. However, compared to the native protein, the yield of globin radical (globin*) was significantly decreased for the Y103F but not the C110A variant Mb upon reaction of the respective proteins with H(2)O(2). In contrast with our expectation that inhibiting pathways of intramolecular electron transfer may lead to enhanced Mb peroxidase activity, mutation of Y103 marginally decreased the rate constant for reaction of Mb with H(2)O(2) (1.4-fold) as judged by stopped-flow kinetic analyses. Consistent with this decrease in rate constant, steady-state analyses of Y103F Mb-derived thioanisole sulfoxidation indicated decreased V(max) and increased K(m) relative to the wild-type control. Additionally, thioanisole sulfoxidation proceeded with lower stereoselectivity, suggesting that Y103 plays a significant role in substrate binding and orientation in the heme pocket of Mb. Together, these results show that electron transfer within the globin portion of the protein is an important modulator of its stability and catalytic activity. Furthermore, the hydrogen-bonding network involving the residues that line the heme pocket of Mb is crucial to both efficient peroxidase activity and stereospecificity.  相似文献   

17.
A cross-linked enzyme aggregate (CLEA) of 3-phytase (EC 3.1.3.8) was synthesised, which was incubated with vanadate and tested as a biocatalyst in the asymmetric sulfoxidation of thioanisole using hydrogen peroxide as the oxidant. The results show that the 3-phytase-CLEA demonstrates a similar efficiency (ca. 95% conversion) and asymmetric induction (ca. 60%) as the free enzyme. Moreover, the 3-phytase-CLEA can be reused at least three times without significant loss of activity. The activity of the 3-phytase in the presence of organic solvents is however still limited. Studies were undertaken to elucidate the role of vanadate on the active site and on the effect of organic solvents on the conformation of the enzyme. The incorporation of vanadate in the active sites of two different phytases could be followed using (51)V NMR and circular dichroism (CD) spectroscopies. (51)V NMR spectra show the incorporation of vanadate into the active site at pH 5.0 and 7.6, and suggest coordination to oxygen functions at two different binding sites, which probably explains the poor enantioselectivity found in the catalytic studies. After addition of H(2)O(2), only at pH 5.0 and with the 3-phytase a V-phytase-peroxide complex could be observed, which is the active species responsible for the oxidation reactions. CD studies showed that the alpha-helical content of the enzyme decreased upon coordination of vanadate, but in the concentration range used in the catalytic studies (<30 microM) the secondary conformation of the enzyme was unchanged. Acetonitrile decreases the alpha-helical content of both phytases from 59% to 51% and from 42% to 34%, in the 3- and 6-phytases, respectively, this being in agreement with the activity loss in the catalytic experiments.  相似文献   

18.
19.
The mechanism of organosulfur oxygenation by peroxidases [lactoperoxidase (LPX), chloroperoxidase, thyroid peroxidase, and horseradish peroxidase] and hydrogen peroxide was investigated by use of para-substituted thiobenzamides and thioanisoles. The rate constants for thiobenzamide oxygenation by LPX/H2O2 were found to correlate with calculated vertical ionization potentials, suggesting rate-limiting single-electron transfer between LPX compound I and the organosulfur substrate. The incorporation of oxygen from 18O-labeled hydrogen peroxide, water, and molecular oxygen into sulfoxides during peroxidase-catalyzed S-oxygenation reactions was determined by LC- and GC-MS. All peroxidases tested catalyzed essentially quantitative oxygen transfer from 18O-labeled hydrogen peroxide into thiobenzamide S-oxide, suggesting that oxygen rebound from the oxoferryl heme is tightly coupled with the initial electron transfer in the active site. Experiments using H2(18)O2, 18O2, and H2(18)O showed that LPX catalyzed approximately 85, 22, and 0% 18O-incorporation into thioanisole sulfoxide oxygen, respectively. These results are consistent with a active site controlled mechanism in which the protein radical form of LPX compound I is an intermediate in LPX-mediated sulfoxidation reactions.  相似文献   

20.
Adediran SA  Pratt RF 《Biochemistry》2008,47(36):9467-9474
All three classes of serine beta-lactamases are inhibited at micromolar levels by 1:1 complexes of catechols with vanadate. Vanadate reacts with catechols at submillimolar concentrations in aqueous buffer at neutral pH in several steps, initially forming 1:1, 1:2, and, possibly, 1:3 complexes. Formation of these complexes is followed by the slower reduction of vanadate (V (V)) to vanadyl (V (IV)) and oxidation of the catechol. Vanadyl-catechol complexes, however, do not inhibit the beta-lactamases. Rate and equilibrium constants of formation of the 1:1 and 1:2 complexes of vanadate with catechol itself and with 2,3-dihydroxynaphthalene were measured by stopped-flow spectrophotometry. Typical examples of all three classes of serine beta-lactamases (the class A TEM-2, class C P99, and class D OXA-1 enzymes) were competitively inhibited by the 1:1 vanadate-catechol complexes. The inhibition was modestly enhanced by hydrophobic substituents on the catechol. The 1:1 vanadate complexes are considerably better inhibitors of the P99 beta-lactamase than 1:1 complexes of catechol with boric acid and are likely to contain penta- or hexacoordinated vanadium rather than tetracooordinated. Molecular modeling showed that a pentacoordinated 1:1 vanadate-catechol complex readily fits into the class C beta-lactamase active site with coordination to the nucleophilic serine hydroxyl oxygen. Such complexes may resemble the pentacoordinated transition states of phosphyl transfer, a reaction also catalyzed by beta-lactamases.  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