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 rate of reaction of ferro- and ferricytochrome c (C(II) and C(III)) with ferri- and ferrocyanide and of C(III) with O2? and CO2? was determined in H2O and in 2H2O in the temperature range 5–35 °C. No isotope effect was evident in any of the reductions of C(III); the apparent energy of activation was identical in H2O and 2H2O. An isotope effect with kH2Ok2H2O = 1.25 to 1.85, depending on pH for instance was observed in the oxidation of C(II), in the slow phase of oxidation which involves conformational changes. An interpretation (supported by evidence from previous work) involving water molecules in the close vicinity of the reaction site on the protein is discussed.  相似文献   

2.
Spectral scans in both the visible (650-450 nm) and the Soret (450-380 nm) regions were recorded for the native enzyme, Compound II, and Compound III of lactoperoxidase and thyroid peroxidase. Compound II for each enzyme (1.7 microM) was prepared by adding a slight excess of H2O2 (6 microM), whereas Compound III was prepared by adding a large excess of H2O2 (200 microM). After these compounds had been formed it was observed that they were slowly reconverted to the native enzyme in the absence of exogenous donors. The pathway of Compound III back to the native enzyme involved Compound II as an intermediate. Reconversion of Compound III to native enzyme was accompanied by the disappearance of H2O2 and generation of O2, with approximately 1 mol of O2 formed for each 2 mol of H2O2 that disappeared. A scheme is proposed to explain these observations, involving intermediate formation of the ferrous enzyme. According to the scheme, Compound III participates in a reaction cycle that effectively converts H2O2 to O2. Iodide markedly affected the interconversions between native enzyme, Compound II, and Compound III for lactoperoxidase and thyroid peroxidase. A low concentration of iodide (4 microM) completely blocked the formation of Compound II when lactoperoxidase or thyroid peroxidase was treated with 6 microM H2O2. When the enzymes were treated with 200 microM H2O2, the same low concentration of iodide completely blocked the formation of Compound III and largely prevented the enzyme degradation that otherwise occurred in the absence of iodide. These effects of iodide are readily explained by (i) the two-electron oxidation of iodide to hypoiodite by Compound I, which bypasses Compound II as an intermediate, and (ii) the rapid oxidation of H2O2 to O2 by the hypoiodite formed in the reaction between Compound I and iodide.  相似文献   

3.
The isotope effect at C-1 on the H2O2-catalysed decarboxylation of pyruvate (used as a model reaction for the enzymic reaction) increases between pH 3 and 10 from 1.0007 +/- 0.0004 to 1.0283 +/- 0.0014 (25 degrees C). This result indicates a change in the rate-determining step from formation of the tetrahedral intermediate to decarboxylation of this intermediate. Practically no isotope fractionation at C-1 (1.0011 +/- 0.0002, pH 6.0, 25 degrees C) is found in the lactate oxidase-catalysed decarboxylation of lactate, which is indicative for the existence of an irreversible O2-dependent step prior to the enzyme-catalysed decarboxylation. In addition, the result provides further evidence that dissociation of pyruvate and H2O2 from the enzyme can be excluded. The isotope effect at C-2 of lactate in the enzymic reaction (1.0048 +/- 0.0004) is attributed to the hydrogen transfer step from lactate to the coenzyme.  相似文献   

4.
An approach has been suggested to study the H/D isotope effect on protein-water and protein-protein intermolecular interactions by determining the content of non-freezing water using low-temperature (1)H NMR in mixed (H2O/D2O) water solutions. Direct data are obtained on the amount of H2O adsorbed (absolute hydration) in presence of the heavy isotope (deuterium D), and isothermals of H2O/D2O fractionation at protein surface groups are presented for temperatures between -10 degrees C and -35 degrees C and solutions of varying composition. The fractionation factor, phi = [x/(1 - x)]/[x(0)/(1 - x(0))], where x and x(0) are the fractions of deuterons in hydration and bulk water, respectively, appeared to be extremely high: phi > 1 at 0.03 < x(0) < 0.10. The high values of phi indicate a decrease in apparent hydration of protein molecules. A probable reason of the effect can be an inter-protein molecular solvent-mediated interaction induced by D2O. The excess of phi over 1 appears to provide a quantitative estimate of the fraction of hydration water affected by such interaction.  相似文献   

5.
A K Mishra  M H Klapper 《Biochemistry》1986,25(23):7328-7336
We have measured, by permeable membrane/mass spectrometry, the 16O/18O, 12C/13C, and solvent H2O/D2O kinetic isotope effects (kie) associated with acyl-alpha-chymotrypsin hydrolysis and transesterification. The hydrolysis of alpha-chymotrypsinyl 2-furoate has a 12C/13C kie of approximately 1.06. Transesterification of the same acyl enzyme shows 16O/18O, 12C/13C, and solvent H2O/D2O kinetic isotope effects of 1.015 (0.003), 1.01-1.02, and 2.226 (0.007), respectively. From the temperature independence of the 16O/18O transesterification kinetic isotope effect and kinetic data reported elsewhere [Wang, C.-L. A., Calvo, K. C., & Klapper, M. H. (1981) Biochemistry 20, 1401-1408], we conclude that there are two active forms of acylchymotrypsin. We also propose that formation of the tetrahedral intermediate is the rate-limiting step in both hydrolysis and transesterification and that the position of the transition state in the transesterification is closer to the starting enzyme ester while that for the hydrolytic reaction is closer to the tetrahedral intermediate. These results are discussed in terms of reaction mechanism plasticity.  相似文献   

6.
Bacterioferritin (EcBFR) of Escherichia coli is an iron-mineralizing hemoprotein composed of 24 identical subunits, each containing a dinuclear metal-binding site known as the "ferroxidase center." The chemistry of Fe(II) binding and oxidation and Fe(III) hydrolysis using H(2)O(2) as oxidant was studied by electrode oximetry, pH-stat, UV-visible spectrophotometry, and electron paramagnetic resonance spin trapping experiments. Absorption spectroscopy data demonstrate the oxidation of two Fe(II) per H(2)O(2) at the ferroxidase center, thus avoiding hydroxyl radical production via Fenton chemistry. The oxidation reaction with H(2)O(2) corresponds to [Fe(II)(2)-P](Z) + H(2)O(2) --> [Fe(III)(2)O-P](Z) + H(2)O, where [Fe(II)(2)-P](Z) represents a diferrous ferroxidase center complex of the protein P with net charge Z and [Fe(III)(2)O-P](Z) a micro-oxo-bridged diferric ferroxidase complex. The mineralization reaction is given by 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2FeOOH((core)) + 4H(+), where two Fe(II) are again oxidized by one H(2)O(2). Hydrogen peroxide is shown to be an intermediate product of dioxygen reduction when O(2) is used as the oxidant in both the ferroxidation and mineralization reactions. Most of the H(2)O(2) produced from O(2) is rapidly consumed in a subsequent ferroxidase reaction with Fe(II) to produce H(2)O. EPR spin trapping experiments show that the presence of EcBFR greatly attenuates the production of hydroxyl radical during Fe(II) oxidation by H(2)O(2), consistent with the ability of the bacterioferritin to facilitate the pairwise oxidation of Fe(II) by H(2)O(2), thus avoiding odd electron reduction products of oxygen and therefore oxidative damage to the protein and cellular components through oxygen radical chemistry.  相似文献   

7.
Treatment of the Cu(II)-Fe(III) derivative of pig allantoic fluid acid phosphatase with hydrogen peroxide caused irreversible inactivation of the enzyme and loss of half of the intensity of the visible absorption spectrum. Phosphate, a competitive inhibitor, protected against this inactivation, suggesting that it occurred as a result of a reaction at the active site. The native Fe(II)-Fe(III) enzyme was irreversibly inactivated by H2O2 to a much smaller extent than the Cu(II)-Fe(III) derivative, whereas the Zn(II)-Fe(III) derivative was stable to H2O2 treatment. The rates of inactivation of the Cu(II)-Fe(III) and Fe(II)-Fe(III) enzymes in the presence of H2O2 were increased by addition of ascorbate. These results suggest involvement of a Fenton-type reaction, generating hydroxyl radicals which react with essential active site groups. Experiments carried out on the Fe(II)-Fe(III) enzyme showed that irreversible inactivation by H2O2 in the presence of ascorbate obeyed pseudo first-order kinetics. A plot of kobs for this reaction against H2O2 concentration (at saturating ascorbate) was hyperbolic, giving kobs(max) = 0.41 +/- 0.025 min-1 and S0.5(H2O2) = 1.16 +/- 0.18 mM. A kinetic scheme is presented to describe the irreversible inactivation, involving hydroxyl radical generation by reaction of H2O2 with Fe(II)-Fe(III) enzyme, reduction of the product Fe(III)-Fe(III) enzyme by ascorbate and reaction of hydroxyl radical with an essential group in the enzyme.  相似文献   

8.
myo-Inositol oxygenase (MIOX) catalyzes the ring-cleaving, four-electron oxidation of its cyclohexan-(1,2,3,4,5,6-hexa)-ol substrate (myo-inositol, MI) to d-glucuronate (DG). The preceding paper [Xing, G., Hoffart, L. M., Diao, Y., Prabhu, K. S., Arner, R. J., Reddy, C. C., Krebs, C., and Bollinger, J. M., Jr. (2006) Biochemistry 45, 5393-5401] demonstrates by M?ssbauer and electron paramagnetic resonance (EPR) spectroscopies that MIOX can contain a non-heme dinuclear iron cluster, which, in its mixed-valent (II/III) and fully oxidized (III/III) states, is perturbed by binding of MI in a manner consistent with direct coordination. In the study presented here, the redox form of the enzyme that activates O(2) has been identified. l-Cysteine, which was previously reported to accelerate turnover, reduces the fully oxidized enzyme to the mixed-valent form, and O(2), the cosubstrate, oxidizes the fully reduced form to the mixed-valent form with a stoichiometry of one per O(2). Both observations implicate the mixed-valent, diiron(II/III) form of the enzyme as the active state. Stopped-flow absorption and freeze-quench EPR data from the reaction of the substrate complex of mixed-valent MIOX [MIOX(II/III).MI] with limiting O(2) in the presence of excess, saturating MI reveal the following cycle: (1) MIOX(II/III).MI reacts rapidly with O(2) to generate an intermediate (H) with a rhombic, g < 2 EPR spectrum; (2) a form of the enzyme with the same absorption features as MIOX(II/III) develops as H decays, suggesting that turnover has occurred; and (3) the starting MIOX(II/III).MI complex is then quantitatively regenerated. This cycle is fast enough to account for the catalytic rate. The DG/O(2) stoichiometry in the reaction, 0.8 +/- 0.1, is similar to the theoretical value of 1, whereas significantly less product is formed in the corresponding reaction of the fully reduced enzyme with limiting O(2). The DG/O(2) yield in the latter reaction decreases as the enzyme concentration is increased, consistent with the hypothesis that initial conversion of the reduced enzyme to the MIOX(II/III).MI complex and subsequent turnover by the mixed-valent form is responsible for the product in this case. The use of the mixed-valent, diiron(II/III) cluster by MIOX represents a significant departure from the mechanisms of other known diiron oxygenases, which all involve activation of O(2) from the II/II manifold.  相似文献   

9.
A severe inactivation is obtained upon the addition of H(2)O(2) to bovine heart F(1)ATPase samples containing Fe(III) in the nucleotide-independent site, and Fe(II) in the ATP-dependent site. EPR spectra at 4.9 K of these samples indicate that H(2)O(2) produces the complete oxidation of Fe(II) to Fe(III) and the concomitant appearance of two protein-derived radical species. The two signals (g = 2.036 and g = 2.007) display a different temperature dependence and saturation behavior. The relaxation properties of the radical at g = 2.036 suggest magnetic interaction with one of the two iron centers. Such events are not observed when H(2)O(2) is added either to native F(1)ATPase containing a high amount of Fe(II) and low amount of Fe(III) or to F(1)ATPase deprived of endogenous Fe and subsequently loaded with only Fe(III) in both sites. It is hypothesized that in F(1)ATPase samples containing both Fe(III) and Fe(II), intramolecular long-range electron transfer may occur from Fe(II) to a high oxidation state species of Fe formed in the nucleotide-independent site upon oxidation of Fe(III) by H(2)O(2).  相似文献   

10.
NADH oxidation by manganese peroxidase (MnP) was done in a reaction mixture including either alpha-hydroxy acid or acetate. The oxidation in the former reaction mixture was inhibited by a catalase and was accelerated by exogenous H2O2, while the oxidation in the latter reaction mixture was inhibited by a superoxide dismutase and was not accelerated by the exogenous H2O2. These results indicated that there are significant differences between the two reaction systems, particularly, in the active oxygen species involved in the reactions. Additionally, the experiment of MnP reduction with Mn(II) suggests that MnP has a separate catalytic activity other than an oxidation of Mn(II) to Mn(III) in the reaction mixture including acetate.  相似文献   

11.
The soluble NAD+-reducing Ni-Fe hydrogenase (SH) from Ralstonia eutropha H16 is remarkable because it cleaves hydrogen in the presence of dioxygen at a unique Ni-Fe active site (Burgdorf et al. (2005) J. Am. Chem. Soc. 127, 576). By X-ray absorption (XAS), FTIR, and EPR spectroscopy, we monitored the structure and oxidation state of its metal centers during H2 turnover. In NADH-activated protein, a change occurred from the (CN)O2Ni(II)(mu-S)2Fe(II)(CN)3(CO) site dominant in the wild-type SH to a standard-like S2Ni(II)(mu-S)2Fe(II)(CN)2(CO) site as the prevailing species in a specific mutant protein, HoxH-H16L. The wild-type SH primarily was active in H2 cleavage. The nonstandard reaction mechanism does not involve stable EPR-detectable trivalent Ni oxidation states, namely, the Ni-A,B,C states as observed in standard hydrogenases. In the HoxH-mutant protein H16L, H2 oxidation was impaired, but H2 production occurred via a stable Ni-C state (Ni(III)-H(-)-Fe(II)), suggesting a reaction sequence similar to that of standard hydrogenases. It is proposed that reductive activation by NADH of both wild-type and H16L proteins causes the release of an oxygen species from Ni and is initiated by electron transfer from a [2Fe-2S] cluster in the HoxU subunit that at first becomes reduced by electrons from NADH. Electrons derived from H2 cleavage, on the other hand, are transferred to NAD+ via a different pathway involving a [4Fe-4S] cluster in HoxY, which is reducible only in wild-type SH but not in the H16L variant.  相似文献   

12.
The catalase reaction has been studied in detail by using myoglobin (Mb) mutants. Compound I of Mb mutants (Mb-I), a ferryl species (Fe(IV)=O) paired with a porphyrin radical cation, is readily prepared by the reaction with a nearly stoichiometric amount of m-chloroperbenzoic acid. Upon the addition of H2O2 to an Mb-I solution, Mb-I is reduced back to the ferric state without forming any intermediates. This indicates that Mb-I is capable of performing two-electron oxidation of H2O2 (catalatic reaction). Gas chromatography-mass spectroscopy analysis of the evolved O2 from a 50:50 mixture of H2(18)O2/H2(16)O2 solution containing H64D or F43H/H64L Mb showed the formation of 18O2 (m/e = 36) and 16O2 (m/e = 32) but not 16O18O (m/e = 34). This implies that O2 is formed by two-electron oxidation of H2O2 without breaking the O-O bond. Deuterium isotope effects on the catalatic reactions of Mb mutants and catalase suggest that the catalatic reactions of Micrococcus lysodeikticus catalase and F43H/H64L Mb proceed via an ionic mechanism with a small isotope effect of less than 4.0, since the distal histidine residue is located at a proper position to act as a general acid-base catalyst for the ionic reaction. In contrast, other Mb mutants such as H64X (X is Ala, Ser, and Asp) and L29H/H64L Mb oxidize H2O2 via a radical mechanism in which a hydrogen atom is abstracted by Mb-I with a large isotope effect in a range of 10-29, due to a lack of the general acid-base catalyst.  相似文献   

13.
Jiang W  Xie J  Nørgaard H  Bollinger JM  Krebs C 《Biochemistry》2008,47(15):4477-4483
We recently showed that the class Ic ribonucleotide reductase (RNR) from the human pathogen Chlamydia trachomatis ( Ct) uses a Mn (IV)/Fe (III) cofactor in its R2 subunit to initiate catalysis [Jiang, W., Yun, D., Saleh, L., Barr, E. W., Xing, G., Hoffart, L. M., Maslak, M.-A., Krebs, C., and Bollinger, J. M., Jr. (2007) Science 316, 1188-1191]. The Mn (IV) site of the novel cofactor functionally replaces the tyrosyl radical used by conventional class I RNRs to initiate substrate radical production. As a first step in evaluating the hypothesis that the use of the alternative cofactor could make the RNR more robust to reactive oxygen and nitrogen species [RO(N)S] produced by the host's immune system [H?gbom, M., Stenmark, P., Voevodskaya, N., McClarty, G., Gr?slund, A., and Nordlund, P. (2004) Science 305, 245-248], we have examined the reactivities of three stable redox states of the Mn/Fe cluster (Mn (II)/Fe (II), Mn (III)/Fe (III), and Mn (IV)/Fe (III)) toward hydrogen peroxide. Not only is the activity of the Mn (IV)/Fe (III)-R2 intermediate stable to prolonged (>1 h) incubations with as much as 5 mM H 2O 2, but both the fully reduced (Mn (II)/Fe (II)) and one-electron-reduced (Mn (III)/Fe (III)) forms of the protein are also efficiently activated by H 2O 2. The Mn (III)/Fe (III)-R2 species reacts with a second-order rate constant of 8 +/- 1 M (-1) s (-1) to yield the Mn (IV)/Fe (IV)-R2 intermediate previously observed in the reaction of Mn (II)/Fe (II)-R2 with O 2 [Jiang, W., Hoffart, L. M., Krebs, C., and Bollinger, J. M., Jr. (2007) Biochemistry 46, 8709-8716]. As previously observed, the intermediate decays by reduction of the Fe site to the active Mn (IV)/Fe (III)-R2 complex. The reaction of the Mn (II)/Fe (II)-R2 species with H 2O 2 proceeds in three resolved steps: sequential oxidation to Mn (III)/Fe (III)-R2 ( k = 1.7 +/- 0.3 mM (-1) s (-1)) and Mn (IV)/Fe (IV)-R2, followed by decay of the intermediate to the active Mn (IV)/Fe (III)-R2 product. The efficient reaction of both reduced forms with H 2O 2 contrasts with previous observations on the conventional class I RNR from Escherichia coli, which is efficiently converted from the fully reduced (Fe 2 (II/II)) to the "met" (Fe 2 (III/III)) form [Gerez, C., and Fontecave, M. (1992) Biochemistry 31, 780-786] but is then only very inefficiently converted from the met to the active (Fe 2 (III/III)-Y (*)) form [Sahlin, M., Sj?berg, B.-M., Backes, G., Loehr, T., and Sanders-Loehr, J. (1990) Biochem. Biophys. Res. Commun. 167, 813-818].  相似文献   

14.
Myeloperoxidase catalyses the conversion of H2O2 and Cl- to hypochlorous acid (HOCl). It also reacts with O2- to form the oxy adduct (compound III). To determine how O2- affects the formation of HOCl, chlorination of monochlorodimedon by myeloperoxidase was investigated using xanthine oxidase and hypoxanthine as a source of O2- and H2O2. Myeloperoxidase was mostly converted to compound III, and H2O2 was essential for chlorination. At pH 5.4, superoxide dismutase (SOD) enhanced chlorination and prevented formation of compound III. However, at pH 7.8, SOD inhibited chlorination and promoted formation of the ferrous peroxide adduct (compound II) instead of compound III. We present spectral evidence for a direct reaction between compound III and H2O2 to form compound II, and for the reduction of compound II by O2- to regenerate native myeloperoxidase. These reactions enable compound III and compound II to participate in the chlorination reaction. Myeloperoxidase catalytically inhibited O2- -dependent reduction of Nitro Blue Tetrazolium. This inhibition is explained by myeloperoxidase undergoing a cycle of reactions with O2-, H2O2 and O2-, with compounds III and II as intermediates, i.e., by myeloperoxidase acting as a combined SOD/catalase enzyme. By preventing the accumulation of inactive compound II, O2- enhances the activity of myeloperoxidase. We propose that, under physiological conditions, this optimizes the production of HOCl and may potentiate oxidant damage by stimulated neutrophils.  相似文献   

15.
During the oxidation of NADH by horseradish peroxidase (HRP-Fe(3+)), superoxide (O(-)(2)) is produced, and HRP-Fe(3+) is converted to compound III. Superoxide dismutase inhibited both the generation of O(-)(2) and the formation of compound III. In contrast, catalase inhibited only the generation of O(-)(2). Under anaerobic conditions, the formation of compound III did not occur in the presence of NADH, thus indicating that compound III is produced via formation of a ternary complex consisting of HRP-Fe(3+), NADH and oxygen. The generation of hydroxyl radicals was dependent upon O(-)(2) and H(2)O(2) produced by HRP-Fe(3+)-NADH. The reaction of compound III with H(2)O(2) caused the formation of compound II without generation of hydroxyl radicals. Only HRP-Fe(3+)-NADH (but not K(+)O(-)(2) and xanthine oxidase-hypoxanthine) was able to induce the conversion of metmyoglobin to oxymyoglobin, thus suggesting the participation of a ternary complex made up of HRP-Fe(2+…)O(2)(…)NAD(.) (but not free O(-)(2) or H(2)O(2)) in the conversion of metmyoglobin to oxymyoglobin. It appears that a cyclic pathway is formed between HRP-Fe(3+), compound III and compound II in the presence of NADH under aerobic conditions, and a ternary complex plays the central roles in the generation of O(-)(2) and hydroxyl radicals.  相似文献   

16.
To clarify the significance of catalase in peroxisomes, we have examined the effect of aminotriazole treatment of rats on the activity of beta-hydroxybutyryl-CoA dehydrogenase in liver peroxisomes. When the effect of H2O2 on the dehydrogenase activity was examined using an extract of liver peroxisomes from aminotriazole-treated rats, the acetoacetyl-CoA-dependent oxidation of NADH was found to increase considerably on the addition of dilute H2O2. Such an effect of H2O2 was not seen on the beta-hydroxybutyryl-CoA-dependent reduction of NAD nor with extracts from untreated animals. We then noticed that similar NADH oxidation was caused non-enzymatically by a mixture of acetoacetyl-CoA and H2O2. The oxidation was dependent on both acetoacetyl-CoA and H2O2, and was blocked by scavengers of oxyradicals such as ascorbate and ethanol. Degradation products formed during the reaction of acetoacetyl-CoA with H2O2 had no NADH oxidizing activity, indicating that effective oxidant(s) were generated during the reaction of H2O2 with acetoacetyl-CoA. No other fatty acyl-CoA so far examined nor acetoacetate could replace acetoacetyl-CoA in this reaction. Therefore, if H2O2 were to be accumulated in peroxisomes, it would decrease both NADH and acetoacetyl-CoA, thus affecting the fatty acyl-CoA beta-oxidation system. These results, together with our previous finding that peroxisomal thiolase was significantly inactivated by H2O2 [Hashimoto, F. & Hayashi, H. (1987) Biochim. Biophys. Acta 921, 142-150] suggest that the role of catalase in peroxisomes is at least in part to protect the fatty acyl-CoA beta-oxidation system from the deleterious action of H2O2.  相似文献   

17.
Hemopexin (HPX), serving as scavenger and transporter of toxic plasma heme, has been postulated to play a key role in the homeostasis of NO. Here, kinetics of HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO are reported. NO reacts reversibly with HPX-heme(II) yielding HPX-heme(II)-NO, according to the minimum reaction scheme: HPX-heme(II)+NO kon<-->koff HPX-heme(II)-NO values of kon, koff, and K (=kon/koff) are (6.3+/-0.3)x10(3)M-1s-1, (9.1+/-0.4)x10(-4)s-1, and (6.9+/-0.6)x10(6)M-1, respectively, at pH 7.0 and 10.0 degrees C. O2 reacts with HPX-heme(II)-NO yielding HPX-heme(III) and NO3-, by means of the ferric heme-bound peroxynitrite intermediate (HPX-heme(III)-N(O)OO), according to the minimum reaction scheme: HPX-heme(II)-NO+O2 hon<--> HPX-heme(III)-N(O)OO l-->HPX-heme(III)+NO3- the backward reaction rate is negligible. Values of hon and l are (2.4+/-0.3)x10(1)M-1s-1 and (1.4+/-0.2)x10(-3)s-1, respectively, at pH 7.0 and 10.0 degrees C. The decay of HPX-heme(III)-N(O)OO (i.e., l) is rate limiting. The HPX-heme(III)-N(O)OO intermediate has been characterized by optical absorption spectroscopy in the Soret region (lambdamax=409 nm and epsilon409=1.51x10(5)M-1cm-1). These results, representing the first kinetic evidence for HPX-heme(II) nitrosylation and O2-mediated oxidation of HPX-heme(II)-NO, might be predictive of transient (pseudo-enzymatic) function(s) of heme carriers.  相似文献   

18.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  相似文献   

19.
The DNA-binding proteins from starved cells (Dps) are a family of proteins induced in microorganisms by oxidative or nutritional stress. Escherichia coli Dps, a structural analog of the 12-subunit Listeria innocua ferritin, binds and protects DNA against oxidative damage mediated by H(2)O(2). Dps is shown to be a Fe-binding and storage protein where Fe(II) oxidation is most effectively accomplished by H(2)O(2) rather than by O(2) as in ferritins. Two Fe(2+) ions bind at each of the 12 putative dinuclear ferroxidase sites (P(Z)) in the protein according to the equation, 2Fe(2+) + P(Z) --> [(Fe(II)(2)-P](FS)(Z+2) + 2H(+). The ferroxidase site (FS) bound iron is then oxidized according to the equation, [(Fe(II)(2)-P](FS)(Z+2) + H(2)O(2) + H(2)O --> [Fe(III)(2)O(2)(OH)-P](FS)(Z-1) + 3H(+), where two Fe(II) are oxidized per H(2)O(2) reduced, thus avoiding hydroxyl radical production through Fenton chemistry. Dps acquires a ferric core of approximately 500 Fe(III) according to the mineralization equation, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(III)OOH((core)) + 4H(+), again with a 2 Fe(II)/H(2)O(2) stoichiometry. The protein forms a similar ferric core with O(2) as the oxidant, albeit at a slower rate. In the absence of H(2)O(2) and O(2), Dps forms a ferrous core of approximately 400 Fe(II) by the reaction Fe(2+) + H(2)O + Cl(-) --> Fe(II)OHCl((core)) + H(+). The ferrous core also undergoes oxidation with a stoichiometry of 2 Fe(II)/H(2)O(2). Spin trapping experiments demonstrate that Dps greatly attenuates hydroxyl radical production during Fe(II) oxidation by H(2)O(2). These results and in vitro DNA damage assays indicate that the protective effect of Dps on DNA most likely is exerted through a dual action, the physical association with DNA and the ability to nullify the toxic combination of Fe(II) and H(2)O(2). In the latter process a hydrous ferric oxide mineral core is produced within the protein, thus avoiding oxidative damage mediated by Fenton chemistry.  相似文献   

20.
Eosinophil recruitment and enhanced nitric oxide (NO) production are characteristic features of asthma and other airway diseases. Eosinophil peroxidase (EPO), a highly cationic hemoprotein secreted by activation of eosinophils, is believed to play a central role in host defense against invading pathogens. The enzyme uses hydrogen peroxide (H2O2) and bromide (Br-), a preferred cosubstrate of EPO, to generate the cytotoxic oxidant hypobromous acid. The aim of this work was to determine whether NO can compete with plasma levels of Br- and steer the enzyme reaction from a 2e- oxidation to a 1e- oxidation pathway. Rapid kinetic measurements were utilized to measure the rate of EPO compounds I and II formation, duration, and decay at 412 and 432 nm, respectively, at 10 degrees C. An EPO-Fe(III) solution supplemented with increasing Br- concentrations was rapidly mixed with fixed amounts of H2O2 in the absence and in the presence of increasing NO concentrations. In the absence of NO, EPO-Fe(III) primarily converted to compound I and, upon H2O2 exhaustion, it decayed rapidly to the ferric form. NO caused a significant increase in the accumulation of EPO compound II, along with a proportional increase in its rate of formation and duration as determined by the time elapsed during catalysis. The time courses for these events have been incorporated into a comprehensive kinetic model. Computer simulations carried out supported the involvement of a conformational intermediate in the EPO compound II complex decay. Collectively, our results demonstrated that NO displays the potential capacity to promote substrate switching by modulating substrate selectivity of EPO.  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