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.
The temperature and pH dependence of the reaction of the superoxide radical anion with ferricytochrome c have been measured using the pulse-radiolysis technique. The temperature dependence of the reaction at low ionic strength yields an activation energy of 31 +/- 5 kJ/mol as compared to 14 +/- 3 kJ/mol for the reaction of CO2.(-) under the same conditions. The pH dependence fits the single pK'a of ferricytochrome c of 9.1. The bimolecular rate constant for the reaction of the superoxide anion with ferricytochrome c at pH 7.8, 21 +/- 2 degrees C, in the presence of 50 mM phosphate and 0.1 mM EDTA is (2.6 +/- 0.1) X 10(5) M-1 s-1. Using this value, 1 unit of superoxide dismutase activity (McCord, J. M., and Fridovich, I. (1969) J. Biol. Chem. 244, 6049-6055) is calculated to be 3.6 +/- 0.3 pmol of enzyme if the assay is performed in a total volume of 3.0 ml. Copper ions reduce the yield of the reaction of ferricytochrome c with CO2.(-). The reactivities of native and singly modified 4-carboxy-2,4-dinitrophenyllysine cytochromes c towards the superoxide anion radical are in the order native greater than 4-carboxy-2,4-dinitrophenyllysine 60 greater than lysine 13 greater than lysine 87 greater than lysine 27 greater than lysine 86 greater than lysine 72, indicating that electron transfer takes place at or close to the solvent accessible heme edge. The mechanism of the reaction is discussed in terms of the approach of superoxide anion radicals to the heme edge and the available molecular orbitals of both heme and free radicals.  相似文献   

2.
Two highly sensitive spectrophotometric methods are developed and described for the measurement of superoxide ion radical derived from KO2 as well as O2*- generated either from the xanthine-xanthine oxidase reaction or by the addition of nicotinamide adenine dinucleotide (NADH) to skeletal muscle sarcoplasmic reticulum (SR) vesicles. These methods allow quantification of superoxide ion concentration by monitoring its reaction with 4-chloro-7-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl), either by recording absorbance of the final reaction product at a wavelength of 470 nm or by measuring its fluorescence emission intensity at 550 nm using an excitation wavelength of 470 nm. The extinction coefficient of the active product was determined to be 4000 M(-1) cm(-1). A lower limit second-order bimolecular rate constant of 1.5+/-0.3x10(5) M(-1) s(-1) was estimated from kinetic stopped-flow analysis for the reaction between NBD-Cl and KO2. A plot of absorbance versus concentration of superoxide was linear over the range 2 to 200 microM KO2, whereas higher sensitivities were obtained from fluorometric measurements down into sub-micromolar concentrations with a limit of detection of 100 nM KO2. This new spectrophotometric assay showed higher specificity when compared with some other commonly used methods for detection of superoxide (e.g., nitroblue tetrazolium). Results presented showed good experimental agreement with rates obtained for the measurement of superoxide ion when compared with other well-known probes such as acetylated ferri cytochrome c and 2,3-bis(2-methoxy-4-nitro-5-sulfophenyl)-5-[(phenylamino)carbonyl]-2H-tetrazolium hydroxide (XTT). A detailed discussion of the advantages and limitations of this new superoxide ion probe is presented.  相似文献   

3.
The ansamycin antibiotic geldanamycin has frequently been used as an inhibitor of heat shock protein 90 (Hsp90), and this agent has been widely employed as a probe to examine the interactions of Hsp90 with endothelial nitric-oxide synthase. Geldanamycin contains a quinone group, which may participate in redox cycling. When geldanamycin was exposed to the flavin-containing enzyme cytochrome P-450 reductase, both semiquinone and superoxide (O(2)(*)(-)) radicals were detected using electron spin resonance. The treatment of endothelial cells with geldanamycin resulted in a dramatic increase in O(2)(*)(-) generation, which was independent of endothelial nitric-oxide synthase, because it was not inhibited by N-nitro-l-arginine methyl ester and also occurred in vascular smooth muscle cells. Diphenylene iodinium inhibited this increase in O(2)(*)(-) by 50%, suggesting that flavin-containing enzymes are involved in geldanamycin-induced O(2)(*)(-) generation. In the absence of cells, geldanamycin directly oxidized ascorbate, consumed oxygen, and produced O(2)(*)(-). Geldanamycin decreased the bioavailable nitric oxide generated by 3,4-dihydrodiazete 1,2-dioxide in smooth muscle cells by 50%, whereas pretreatment with superoxide dismutase inhibited the effect of geldanamycin. These findings demonstrate that geldanamycin generates O(2)(*)(-), which scavenges nitric oxide, leading to loss of its bioavailability. This effect is independent of the inhibition of Hsp90 and indicates that geldanamycin cannot be used as a specific inhibitor of Hsp90. In light of these findings, the studies using geldanamycin as an inhibitor of Hsp90 should be interpreted with caution.  相似文献   

4.
The rate and mechanism of autoxidation of soluble ferrocytochrome b5, prepared from liver microsomal suspensions, appear to reflect an intrinsic property of membrane-bound cytochrome b5. The first-order rate constant for autoxidation of trypsin-cleaved ferrocytochrome b5, prepared by reduction with dithionite, was 2.00 X 10(-3) +/- 0.19 X 10(-3) S-1 (mean +/- S.E.M., n =8) when measured at 30 degrees C in 10 mM-phosphate buffer, pH 7.4. At 37 degrees C in aerated 10 mM-phosphate buffer (pH 7.4)/0.15 M-KCl, the rate constant was 5.6 X 10(-3) S-1. The autoxidation reaction was faster at lower pH values and at high ionic strengths. Unlike ferromyoglobin, the autoxidation reaction of which is maximal at low O2 concentrations, autoxidation of ferrocytochrome b5 showed a simple O2-dependence with an apparent Km for O2 of 2.28 X 10(-4) M (approx. 20kPa or 150mmHg)9 During autoxidation, 0.25 mol of O2 was consumed per mol of cytochrome oxidized. Cyanide, nucleophilic anions, EDTA and catalase each had little or no effect on autoxidation rates. Adrenaline significantly enhanced autoxidation rates, causing a tenfold increase at 0.6 mM. Ferrocytochrome b5 reduced an excess of cytochrome c in a biphasic manner. An initial rapid phase, independent of O2 concentration, was unaffected by superoxide dismutase. A subsequent slower phase, which continued for up to 60 min, was retarded at low O2 concentrations and inhibited by 65% by superoxide dismutase at a concentration of 3 mug/ml. It is concluded that autoxidation is responsible for a significant proportion of electron flow between cytochrome b5 and O2 in liver endoplasmic membranes, this reaction being capable of generating superoxide anions. A biological role for the reaction is discussed.  相似文献   

5.
Oxygen activation at a carboxylate-bridged diiron cluster is employed by a number of enzymes for diverse biological functions. The mechanisms by which O(2) is activated at the diferrous clusters have been studied in detail and peroxodiferric reaction intermediates have been observed in several of these diiron proteins. To understand further the magnetic properties of this common reaction intermediate, we have used M?ssbauer spectroscopy to determine the magnitude and sign of the exchange coupling constant J (in the exchange Hamiltonian J S(1) x S(2)) of the peroxodiferric intermediates generated during the reactions of O(2) with two different proteins, the recombinant M ferritin from frog and the site-directed variant W48F/D84E of the R2 subunit of ribonucleotide reductase from Escherichia coli. Both intermediates are antiferromagnetically coupled with a moderate coupling constant J of 50+/-10 cm(-1) for R2-W48F/D84E and 75+/-10 cm(-1) for M ferritin. This work demonstrates the capability of M?ssbauer spectroscopy to determine exchange coupling constants of diiron complexes, including reaction intermediates. The approach and its limitations are described.  相似文献   

6.
Superoxide reductases are a class of non-haem iron enzymes which catalyse the monovalent reduction of the superoxide anion O2- into hydrogen peroxide and water. Treponema pallidum (Tp), the syphilis spirochete, expresses the gene for a superoxide reductase called neelaredoxin, having the iron protein rubredoxin as the putative electron donor necessary to complete the catalytic cycle. In this work, we present the first cloning, overexpression in Escherichia coli and purification of the Tp rubredoxin. Spectroscopic characterization of this 6 kDa protein allowed us to calculate the molar absorption coefficient of the 490 nm feature of ferric iron, epsilon=6.9+/-0.4 mM(-1) cm(-1). Moreover, the midpoint potential of Tp rubredoxin, determined using a glassy carbon electrode, was -76+/-5 mV. Reduced rubredoxin can be efficiently reoxidized upon addition of Na(2)IrCl(6)-oxidized neelaredoxin, in agreement with a direct electron transfer between the two proteins, with a stoichiometry of the electron transfer reaction of one molecule of oxidized rubredoxin per one molecule of neelaredoxin. In addition, in presence of a steady-state concentration of superoxide anion, the physiological substrate of neelaredoxin, reoxidation of rubredoxin was also observed in presence of catalytic amounts of superoxide reductase, and the rate of rubredoxin reoxidation was shown to be proportional to the concentration of neelaredoxin, in agreement with a bimolecular reaction, with a calculated k(app)=180 min(-1). Interestingly, similar experiments performed with a rubredoxin from the sulfate-reducing bacteria Desulfovibrio vulgaris resulted in a much lower value of k(app)=4.5 min(-1). Altogether, these results demonstrated the existence for a superoxide-mediated electron transfer between rubredoxin and neelaredoxin and confirmed the physiological character of this electron transfer reaction.  相似文献   

7.
Heo J  Campbell SL 《Biochemistry》2006,45(7):2200-2210
Ras GTPases cycle between inactive GDP-bound and active GTP-bound states to modulate a diverse array of processes involved in cellular growth control. We have previously shown that both NO/O(2) (via nitrogen dioxide, (*)NO(2)) and superoxide radical anion (O(2)(*)(-)) promote Ras guanine nucleotide dissociation. We now show that hydrogen peroxide in the presence of transition metals (i.e., H(2)O(2)/transition metals) and peroxynitrite also trigger radical-based Ras guanine nucleotide dissociation. The primary redox-active reaction species derived from H(2)O(2)/transition metals and peroxynitrite is O(2)(*)(-) and (*)NO(2), respectively. A small fraction of hydroxyl radical (OH(*)) is also present in both. We also show that both carbonate radical (CO(3)(*)(-)) and (*)NO(2), derived from the mixture of peroxynitrite and bicarbonate, facilitate Ras guanine nucleotide dissociation. We further demonstrate that NO/O(2) and O(2)(*)(-) promote Ras GDP exchange with GTP in the presence of a radical-quenching agent, ascorbate, or NO, and generation of Ras-GTP promotes high-affinity binding of the Ras-binding domain of Raf-1, a downstream effector of Ras. S-Nitrosylated Ras (Ras-SNO) can be formed when NO serves as a radical-quenching agent, and hydroxyl radical but not (*)NO(2) or O(2)(*)(-) can further react with Ras-SNO to modulate Ras activity in vitro. However, given the lack of redox specificity associated with the high redox potential of OH(*), it is unclear whether this reaction occurs under physiological conditions.  相似文献   

8.
The exact role of superoxide radicals (O(2)(*)(-)) in apoptosis is still a matter of debate. The main objective of the present study is to evaluate the apoptotic signalling pathway initiated by O(2)(*)(-). The reductive reaction of sodium selenite with glutathione was used as the intracellular O(2)(*)(-)-generating system. When cells were exposed to 5 to 25 microM selenite, a temporal pattern of apoptotic events was observed following the elevation of O(2)(*)(-), in which cytochrome c release and mitochondrial depolarization preceded caspase-3 activation and DNA fragmentation. The simultaneous treatment with N-acetylcysteine and 4-hydroxy-2,2,6, 6-tetramethylpiperidine-N-oxyl markedly reduced O(2)(*)(-) level and suppressed the mitochondrial changes and the downstream apoptotic events. Moreover, pretreatment with cyclosporin A plus trifluoperazine, two mitochondrial permeability transition (MPT) inhibitors, was capable of attenuating O(2)(*)(-)-mediated cytochrome c release and mitochondrial depolarization, and subsequently inhibiting apoptosis. Thus, the present results provide convincing evidence that O(2)(*)(-) generated from the reductive reaction of selenite with GSH is capable of triggering a mitochondria-dependent apoptotic pathway. Such knowledge may not only help to obtain a better understanding of the apoptotic effect of selenite per se, but of the role of O(2)(*)(-) in initiation and execution of apoptosis.  相似文献   

9.
Superoxide reductase (SOR) is a metalloenzyme that catalyzes the reduction of O2*- to H2O2 and provides an antioxidant mechanism in some anaerobic and microaerophilic bacteria. Its active site contains an unusual mononuclear ferrous center (center II). Protonation processes are essential for the reaction catalyzed by SOR, since two protons are required for the formation of H2O2. We have investigated the acido-basic and pH dependence of the redox properties of the active site of SOR from Desulfoarculus baarsii, both in the absence and in the presence of O2*-. In the absence of O2*-, the reduction potential and the absorption spectrum of the iron center II exhibit a pH transition. This is consistent with the presence of a base (BH) in close proximity to the iron center which modulates its reduction properties. Studies of mutants of the closest charged residues to the iron center II (E47A and K48I) show that neither of these residues are the base responsible for the pH transitions. However, they both interact with this base and modulate its pKa value. By pulse radiolysis, we confirm that the reaction of SOR with O2*- involves two reaction intermediates that were characterized by their absorption spectra. The precise step of the catalytic cycle in which one protonation takes place was identified. The formation of the first reaction intermediate, from a bimolecular reaction of SOR with O2*-, does not involve proton transfer as a rate-limiting step, since the rate constant k1 does not vary between pH 5 and pH 9.5. On the other hand, the rate constant k2 for the formation of the second reaction intermediate is proportional to the H+ concentration in solution, suggesting that the proton arises directly from the solvent. In fact, BH, E47, and K48 have no role in this step. This is consistent with the first intermediate being an iron(III)-peroxo species and the second one being an iron(III)-hydroperoxo species. We propose that BH may be involved in the second protonation process corresponding to the release of H2O2 from the iron(III)-hydroperoxo species.  相似文献   

10.
Aqueous solutions of osmium tetroxide (OsO4) have been injected into arthritic knees for the past 45 years to chemically destroy diseased tissue, in a procedure termed "chemical synovectomy." Arthritis is an inflammatory disease. The primary inflammatory chemical species are the superoxide anion radical (O2.-) and nitric oxide (.NO), which combine to form the peroxynitrite anion (ONOO-). Here we show that OsO4 does not react with ONOO- but very efficiently catalyzes the dismutation of O2.- to O2 and H2O2. Using the pulse-radiolysis technique, the catalytic rate constant has been determined to be (1.43+/-0.04) x 10(9) M-1 s-1, independent of the pH in the 5.1-8.7 range. This value is about half that for the natural Cu,Zn-superoxide dismutase (Cu,Zn-SOD). Per unit mass, OsO4 is about 60 times more active than Cu,Zn-SOD. The catalytically active couple is OsVIII/OsVII, OsVIII oxidizing O2.- to O2 with a bimolecular rate constant of k=(2.6+/-0.1)x10(9) M-1 s-1 and OsVII reducing it to H2O2 with a bimolecular rate constant of (1.0+/-0.1)x10(9) M-1 s-1. Although lower valent osmium species are intrinsically poor catalysts, they are activated through oxidation by O2.- to the catalytic OsVIII/OsVII redox couple. The OsVIII/OsVII catalyst is stable to biochemicals other than proteins and peptides comprising histidine, cysteine, and dithiols.  相似文献   

11.
Flock U  Watmough NJ  Adelroth P 《Biochemistry》2005,44(31):10711-10719
The respiratory nitric oxide reductase (NOR) from Paracoccus denitrificans catalyzes the two-electron reduction of NO to N(2)O (2NO + 2H(+) + 2e(-) --> N(2)O + H(2)O), which is an obligatory step in the sequential reduction of nitrate to dinitrogen known as denitrification. NOR has four redox-active cofactors, namely, two low-spin hemes c and b, one high-spin heme b(3), and a non-heme iron Fe(B), and belongs to same superfamily as the oxygen-reducing heme-copper oxidases. NOR can also use oxygen as an electron acceptor; this catalytic activity was investigated in this study. We show that the product in the steady-state reduction of oxygen is water. A single turnover of the fully reduced NOR with oxygen was initiated using the flow-flash technique, and the progress of the reaction monitored by time-resolved optical absorption spectroscopy. Two major phases with time constants of 40 micros and 25 ms (pH 7.5, 1 mM O(2)) were observed. The rate constant for the faster process was dependent on the O(2) concentration and is assigned to O(2) binding to heme b(3) at a bimolecular rate constant of 2 x 10(7) M(-)(1) s(-)(1). The second phase (tau = 25 ms) involves oxidation of the low-spin hemes b and c, and is coupled to the uptake of protons from the bulk solution. The rate constant for this phase shows a pH dependence consistent with rate limitation by proton transfer from an internal group with a pK(a) = 6.6. This group is presumably an amino acid residue that is crucial for proton transfer to the catalytic site also during NO reduction.  相似文献   

12.
Boccini F  Herold S 《Biochemistry》2004,43(51):16393-16404
The strong oxidizing and nitrating agent peroxynitrite has been shown to diffuse into erythrocytes and oxidize oxyhemoglobin (oxyHb) to metHb. Because the value of the second-order rate constant for this reaction is on the order of 10(4) M(-)(1) s(-)(1) and the oxyHb concentration is about 20 mM (expressed per heme), this process is rather fast and oxyHb is considered a sink for peroxynitrite. In this work, we showed that the reaction of oxyHb with peroxynitrite, both in the presence and absence of CO(2), proceeds via the formation of oxoiron(iv)hemoglobin (ferrylHb), which in a second step is reduced to metHb and nitrate by its reaction with NO(2)(*). In the presence of physiological relevant amounts of CO(2), ferrylHb is generated by the reaction of NO(2)(*) with the coordinated superoxide of oxyHb (HbFe(III)O(2)(*)(-)). This reaction proceeds via formation of a peroxynitrato-metHb complex (HbFe(III)OONO(2)), which decomposes to generate the one-electron oxidized form of ferrylHb, the oxoiron(iv) form of hemoglobin with a radical localized on the globin. CO(3)(*)(-), the second radical formed from the reaction of peroxynitrite with CO(2), is also scavenged efficiently by oxyHb, in a reaction that finally leads to metHb production. Taken together, our results indicate that oxyHb not only scavenges peroxynitrite but also the radicals produced by its decomposition.  相似文献   

13.
S Loo  J E Erman 《Biochemistry》1975,14(15):3467-3470
The rate of the reaction between cytochrome c peroxidase and hydrogen peroxide was investigated using the stopped-flow technique. The apparent bimolecular rate constant was determined between pH 3.3 and pH 11 as a function of ionic strength. The pH dependence of the apparent bimolecular rate constant can be explained by assuming that two ionizable groups on the enzyme strongly influence the rate of the reaction. At 0.1 M ionic strength, a group with a pKa of 5.5 must be unprotonated and a group with a pKa of 9.8 must be protonated for the enzyme to react rapidly with hydrogen peroxide. The apparent acid dissociation constants depend upon the ionic strength. The true bimolecular rate constant has a value of (4.5 +/- 0.3) X 10(7) M-1 sec-1 and is independent of ionic strength.  相似文献   

14.
We directly measured the absolute reactivity of 17-beta-estradiol (E2) and several phenolic model compounds for E2 toward t-butoxy radical (t-BuO*) by nanosecond time-resolved optical spectroscopy. Compared to other phenols, E2 is a moderate, but not strong deactivator of oxyradicals. The absolute bimolecular rate constant for H-atom transfer from E2 to t-BuO* is 1.3 +/- 0.3 x 10(9) M(-1) x s(-1) (23 degrees C, benzene). We estimate the O-H bond strength of 17-beta-estradiol to be approximately 85 +/- 2 kcal/mol and calculate the reaction rate constant of E2 toward peroxy radical to be 10(5) M(-1) x s(-1) at 37 degrees C. The conjugate phenoxy radical of 17-beta-estradiol, E2O*, is unusually reactive toward alpha-tocopherol and ascorbate by H-atom transfer in homogeneous solution (10(8)-10(9) M(-1) x s(-1)). Our findings suggest that E2 functions in vivo as a highly localized, synergistic biological antioxidant. This may partly explain the clinical effectiveness of ovarian steroids in delaying the manifestations of Alzheimer's Disease as well as in protecting against cardiovascular pathologies. In the absence of complementary antioxidant synergists, E2O* is expected to be a pro-oxidant.  相似文献   

15.
Lactoperoxidase (LPO) is found in mucosal surfaces and exocrine secretions including milk, tears, and saliva and has physiological significance in antimicrobial defense which involves (pseudo-)halide oxidation. LPO compound III (a ferrous-dioxygen complex) is known to be formed rapidly by an excess of hydrogen peroxide and could participate in the observed catalase-like activity of LPO. The present anaerobic stopped-flow kinetic analysis was performed in order to elucidate the catalytic mechanism of LPO and the kinetics of compound III formation by probing the reactivity of ferrous LPO with hydrogen peroxide and molecular oxygen. It is shown that ferrous LPO heterolytically cleaves hydrogen peroxide forming water and oxyferryl LPO (compound II). The two-electron oxidation reaction follows second-order kinetics with the apparent bimolecular rate constant being (7.2+/-0.3) x 10(4) M(-1) s(-1) at pH 7.0 and 25 degrees C. The H2O2-mediated conversion of compound II to compound III follows also second-order kinetics (220 M(-1) s(-1) at pH 7.0 and 25 degrees C). Alternatively, compound III is also formed by dioxygen binding to ferrous LPO at an apparent bimolecular rate constant of (1.8+/-0.2) x 10(5) M(-1) s(-1). Dioxygen binding is reversible and at pH 7.0 the dissociation constant (K(D)) of the oxyferrous form is 6 microM. The rate constant of dioxygen dissociation from compound III is higher than conversion of compound III to ferric LPO, which is not affected by the oxygen concentration and follows a biphasic kinetics. A reaction cycle including the redox intermediates compound II, compound III, and ferrous LPO is proposed, which explains the observed (pseudo-)catalase activity of LPO in the absence of one-electron donors. The relevance of these findings in LPO catalysis is discussed.  相似文献   

16.
Yeh AP  Hu Y  Jenney FE  Adams MW  Rees DC 《Biochemistry》2000,39(10):2499-2508
Superoxide reductase (SOR) is a blue non-heme iron protein that functions in anaerobic microbes as a defense mechanism against reactive oxygen species by catalyzing the reduction of superoxide to hydrogen peroxide [Jenney, F. E., Jr., Verhagen, M. F. J. M., Cui, X. , and Adams, M. W. W. (1999) Science 286, 306-309]. Crystal structures of SOR from the hyperthermophilic archaeon Pyrococcus furiosus have been determined in the oxidized and reduced forms to resolutions of 1.7 and 2.0 A, respectively. SOR forms a homotetramer, with each subunit adopting an immunoglobulin-like beta-barrel fold that coordinates a mononuclear, non-heme iron center. The protein fold and metal center are similar to those observed previously for the homologous protein desulfoferrodoxin from Desulfovibrio desulfuricans [Coelho, A. V., Matias, P., Fül?p, V., Thompson, A., Gonzalez, A., and Carrondo, M. A. (1997) J. Bioinorg. Chem. 2, 680-689]. Each iron is coordinated to imidazole nitrogens of four histidines in a planar arrangement, with a cysteine ligand occupying an axial position normal to this plane. In two of the subunits of the oxidized structure, a glutamate carboxylate serves as the sixth ligand to form an overall six-coordinate, octahedral coordinate environment. In the remaining two subunits, the sixth coordination site is either vacant or occupied by solvent molecules. The iron centers in all four subunits of the reduced structure exhibit pentacoordination. The structures of the oxidized and reduced forms of SOR suggest a mechanism by which superoxide accessibility may be controlled and define a possible binding site for rubredoxin, the likely physiological electron donor to SOR.  相似文献   

17.
Superoxide reductases (SORs) contain a novel square pyramidal ferrous [Fe(NHis)(4)(SCys)] site that rapidly reduces superoxide to hydrogen peroxide. Here we report extensive pulse radiolysis studies on recombinant two-iron SOR (2Fe-SOR) from Desulfovibrio vulgaris. The results support and elaborate on our originally proposed scheme for reaction of the [Fe(NHis)(4)(SCys)] site with superoxide [Coulter, E. D., Emerson, J. E., Kurtz, D. M., Jr., and Cabelli, D. E. (2000) J. Am. Chem. Soc. 122, 11555-11556]. This scheme consists of second-order diffusion-controlled formation of an intermediate absorbing at approximately 600 nm, formulated as a ferric-(hydro)peroxo species, and its decay to the carboxylate-ligated ferric [Fe(NHis)(4)(SCys)] site with loss of hydrogen peroxide. The second-order rate constant for formation of the 600-nm intermediate is essentially pH-independent (pH 5-9.5), shows no D(2)O solvent isotope effect at pH 7.7, and decreases with increasing ionic strength. These data indicate that formation of the intermediate does not involve a rate-determining protonation, and are consistent with interaction of the incoming superoxide anion with a positive charge at or near the ferrous [Fe(NHis)(4)(SCys)] site. The rate constant for decay of the 600-nm intermediate follows the pH-dependent rate law: k(2)(obs) = k(2)'[H(+)] + k(2)' ' and shows a significant D(2)O solvent isotope effect at pH 7.7. The values of k(2)' and k(2)' ' indicate that the 600-nm intermediate decays via diffusion-controlled protonation at acidic pHs and a first-order process involving either water or a water-exchangeable proton on the protein at basic pHs. The formation and decay rate constants for an E47A variant of 2Fe-SOR are not significantly perturbed from their wild-type values, indicating that the conserved glutamate carboxylate does not directly displace the (hydro)peroxo ligand of the intermediate at basic pHs. The kinetics of a K48A variant are consistent with participation of the lysyl side chain in directing the superoxide toward the active site and in directing the protonation pathway of the ferric-(hydro)peroxo intermediate toward release of hydrogen peroxide.  相似文献   

18.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

19.
Copper, zinc superoxide dismutase (CuZnSOD) from bovine erythrocytes and iron superoxide dismutase from Escherichia coli (FeSOD) were immobilized on 3-mercaptopropionic acid (MPA)-modified gold electrodes, respectively. The characterization of the SOD electrodes showed a quasi-reversible, electrochemical redox behavior with a formal potential of 47+/-4 mV and -154+/-5 mV (vs. Ag/AgCl, 1 M KCl) for surface adsorbed CuZnSOD and FeSOD, respectively. The heterogeneous electron transfer rate constants were determined to be about 65 and 35/s, respectively. Covalent fixation of both SODs was also feasible with only slight changes in the formal potential. The interaction of superoxide radicals (O(2)(-)) with the SOD electrode was investigated. No catalytic current could be observed. However, due to the fast cyclic redox reaction of SOD with superoxide, the communication of the protein with the electrode was strongly influenced. The amperometric detection of superoxide radicals is discussed.  相似文献   

20.
Generation of superoxide anion and hydrogen peroxide during enzymatic oxidation of 3-(3,4-dihydroxyphenyl)-DL-alanine (DOPA) has been studied. The ability of DOPA to react with O2*- has been revealed. EPR spectrum of DOPA-semiquinone formed upon oxidation of DOPA by O2*- was observed using spin stabilization technique of ortho-semiquinones by Zn2+ ions. Simultaneously, the oxidation of DOPA by O2*- was found to produce hydrogen peroxide (H2O2). The analysis of H2O2 formation upon oxidation of DOPA by O2*- using 1-hydroxy-3-carboxy-pyrrolidine (CP-H), and SOD as competitive reagents for superoxide provides consistent values of the rate constant for the reaction between DOPA and O2*- being equal to (3.4+/-0.6)x10(5) M(-1) s(-1).The formation of H2O2 during enzymatic oxidation of DOPA by phenoloxidase (PO) has been shown. The H2O2 production was found to be SOD-sensitive. The inhibition of H2O2 production by SOD was about 25% indicating that H2O2 is produced both from superoxide anion and via two-electron reduction of oxygen at the enzyme. The attempts to detect superoxide production during enzymatic oxidation of DOPA using a number of spin traps failed apparently due to high value of the rate constant for DOPA interaction with O2*-.  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