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 156 毫秒
1.
We investigated the potential role of the co-substrate, thiocyanate (SCN-), in modulating the catalytic activity of myeloperoxidase (MPO) and other members of the mammalian peroxidase superfamily (lactoperoxidase (LPO) and eosinophil peroxidase (EPO)). Pre-incubation of SCN- with MPO generates a more complex biological setting, because SCN- serves as either a substrate or inhibitor, causing diverse impacts on the MPO heme iron microenvironment. Consistent with this hypothesis, the relationship between the association rate constant of nitric oxide binding to MPO-Fe(III) as a function of SCN- concentration is bell-shaped, with a trough comparable with normal SCN- plasma levels. Rapid kinetic measurements indicate that MPO, EPO, and LPO Compound I formation occur at rates slower than complex decay, and its formation serves to simultaneously catalyze SCN- via 1e- and 2e- oxidation pathways. For the three enzymes, Compound II formation is a fundamental feature of catalysis and allows the enzymes to operate at a fraction of their possible maximum activities. MPO and EPO Compound II is relatively stable and decays gradually within minutes to ground state upon H2O2 exhaustion. In contrast, LPO Compound II is unstable and decays within seconds to ground state, suggesting that SCN- may serve as a substrate for Compound II. Compound II formation can be partially or completely prevented by increasing SCN- concentration, depending on the experimental conditions. Collectively, these results illustrate for the first time the potential mechanistic differences of these three enzymes. A modified kinetic model, which incorporates our current findings with the mammalian peroxidases classic cycle, is presented.  相似文献   

2.
The manganese peroxidase (MnP), from the lignin-degrading fungus Phanerochaete chrysosporium, an H2O2-dependent heme enzyme, oxidizes a variety of organic compounds but only in the presence of Mn(II). The homogeneous enzyme rapidly oxidizes Mn(II) to Mn(III) with a pH optimum of 5.0; the latter was detected by the characteristic spectrum of its lactate complex. In the presence of H2O2 the enzyme oxidizes Mn(II) significantly faster than it oxidizes all other substrates. Addition of 1 M equivalent of H2O2 to the native enzyme in 20 mM Na-succinate, pH 4.5, yields MnP compound II, characterized by a Soret maximum at 416 nm. Subsequent addition of 1 M equivalent of Mn(II) to the compound II form of the enzyme results in its rapid reduction to the native Fe3+ species. Mn(III)-lactate oxidizes all of the compounds which are oxidized by the enzymatic system. The relative rates of oxidation of various substrates by the enzymatic and chemical systems are similar. In addition, when separated from the polymeric dye Poly B by a semipermeable membrane, the enzyme in the presence of Mn(II)-lactate and H2O2 oxidizes the substrate. All of these results indicate that the enzyme oxidizes Mn(II) to Mn(III) and that the Mn(III) complexed to lactate or other alpha-hydroxy acids acts as an obligatory oxidation intermediate in the oxidation of various dyes and lignin model compounds. In the absence of exogenous H2O2, the Mn-peroxidase oxidized NADH to NAD+, generating H2O2 in the process. The H2O2 generated by the oxidation of NADH could be utilized by the enzyme to oxidize a variety of other substrates.  相似文献   

3.
The present study characterizes the serial reactions of H2O2 with compounds I and II of lignin peroxidase isozyme H1. These two reactions constitute part of the pathway leading to formation of the oxy complex (compound III) from the ferric enzyme. Compounds II and III are the only complexes observed; no compound III* is observed. Compound III* is proposed to be an adduct of compound III with H2O2, formed from the complexation of compound III with H2O2 (Wariishi, H., and Gold, M. H. (1990) J. Biol. Chem. 265, 2070-2077). We provide evidence that demonstrates that the spectral data, on which the formation of compound III* is based, are merely an artifact caused by enzyme instability and, therefore, rule out the existence of compound III*. The reactions of compounds II and III with H2O2 are pH-dependent, similar to that observed for reactions of compounds I and II with the reducing substrate veratryl alcohol. The spontaneous decay of the compound III of lignin peroxidase results in the reduction of ferric cytochrome c. The reduction is inhibited by superoxide dismutase, indicating that superoxide is released during the decay. Therefore, the lignin peroxidase compound III decays to the ferric enzyme through the dissociation of superoxide. This mechanism is identical with that observed with oxymyoglobin and oxyhemoglobin but different from that for horseradish peroxidase. Compound III is capable of reacting with small molecules, such as tetranitromethane (a superoxide scavenger) and fluoride (a ligand for the ferric enzyme), resulting in ferric enzyme and fluoride complex formation, respectively.  相似文献   

4.
Wu W  Chen Y  d'Avignon A  Hazen SL 《Biochemistry》1999,38(12):3538-3548
Detection of specific reaction products is a powerful approach for dissecting out pathways that mediate oxidative damage in vivo. Eosinophil peroxidase (EPO), an abundant protein secreted from activated eosinophils, has been implicated in promoting oxidative tissue injury in conditions such as asthma, allergic inflammatory disorders, cancer, and helminthic infections. This unique heme protein amplifies the oxidizing potential of H2O2 by utilizing plasma levels of Br- as a cosubstrate to form potent brominating agents. Brominated products might thus serve as powerful tools for identifying sites of eosinophil-mediated tissue injury in vivo; however, structural identification and characterization of specific brominated products formed during EPO-catalyzed oxidation have not yet been reported. Here we explore the role of EPO and myeloperoxidase (MPO), a related leukocyte protein, in promoting protein oxidative damage through bromination and demonstrate that protein tyrosine residues serve as endogenous traps of reactive brominating species forming stable ring-brominated adducts. Exposure of the amino acid L-tyrosine to EPO, H2O2, and physiological concentrations of halides (100 mM Cl-, 相似文献   

5.
The reactions of lactoperoxidase (LPO) intermediates compound I, compound II and compound III, with nitrite (NO2(-)) were investigated. Reduction of compound I by NO2(-) was rapid (k2 = 2.3 x 10(7) M(-1) x s(-1); pH = 7.2) and compound II was not an intermediate, indicating that NO2* radicals are not produced when NO2(-) reacts with compound I. The second-order rate constant for the reaction of compound II with NO2(-) at pH = 7.2 was 3.5 x 10(5) M(-1) x s(-1). The reaction of compound III with NO2(-) exhibited saturation behaviour when the observed pseudo first-order rate constants were plotted against NO2(-) concentrations and could be quantitatively explained by the formation of a 1 : 1 ratio compound III/NO2(-) complex. The Km of compound III for NO2(-) was 1.7 x 10(-4) M and the first-order decay constant of the compound III/ NO2(-) complex was 12.5 +/- 0.6 s(-1). The second-order rate constant for the reaction of the complex with NO2(-) was 3.3 x 10(3) M(-1) x s(-1). Rate enhancement by NO2(-) does not require NO2* as a redox intermediate. NO2(-) accelerates the overall rate of catalysis by reducing compound II to the ferric state. With increasing levels of H2O2, there is an increased tendency for the catalytically dead-end intermediate compound III to form. Under these conditions, the 'rescue' reaction of NO2(-) with compound III to form compound II will maintain the peroxidatic cycle of the enzyme.  相似文献   

6.
Berka V  Yeh HC  Gao D  Kiran F  Tsai AL 《Biochemistry》2004,43(41):13137-13148
Tetrahydrobiopterin (BH(4)), not dihydrobiopterin or biopterin, is a critical element required for NO formation by nitric oxide synthase (NOS). To elucidate how BH(4) affects eNOS activity, we have investigated BH(4) redox functions in the endothelial NOS (eNOS). Redox-state changes of BH(4) in eNOS were examined by chemical quench/HPLC analysis during the autoinactivation of eNOS using oxyhemoglobin oxidation assay for NO formation at room temperature. Loss of NO formation activity linearly correlated with BH(4) oxidation, and was recovered by overnight incubation with fresh BH(4). Thus, thiol reagents commonly added to NOS enzyme preparations, such as dithiothreitol and beta-mercaptoethanol, probably preserve enzyme activity by preventing BH(4) oxidation. It has been shown that conversion of L-arginine to N-hydroxy-L-arginine in the first step of NOS catalysis requires two reducing equivalents. The first electron that reduces ferric to the ferrous heme is derived from flavin oxidation. The issue of whether BH(4) supplies the second reducing equivalent in the monooxygenation of eNOS was investigated by rapid-scan stopped-flow and rapid-freeze-quench EPR kinetic measurements. In the presence of L-arginine, oxygen binding kinetics to ferrous eNOS or to the ferrous eNOS oxygenase domain (eNOS(ox)) followed a sequential mechanism: Fe(II) <--> Fe(II)O(2) --> Fe(III) + O(2)(-). Without L-arginine, little accumulation of the Fe(II)O(2) intermediate occurred and essentially a direct optical transition from the Fe(II) form to the Fe(III) form was observed. Stabilization of the Fe(II)O(2) intermediate by L-arginine has been established convincingly. On the other hand, BH(4) did not have significant effects on the oxygen binding and decay of the oxyferrous intermediate of the eNOS or eNOS oxygenase domain. Rapid-freeze-quench EPR kinetic measurements in the presence of L-arginine showed a direct correlation between BH(4) radical formation and decay of the Fe(II)O(2) intermediate, indicating that BH(4) indeed supplies the second electron for L-arginine monooxygenation in eNOS.  相似文献   

7.
Eosinophil peroxidase (EPO) has been implicated in promoting oxidative tissue injury in conditions ranging from asthma and other allergic inflammatory disorders to cancer and parasitic/helminthic infections. Studies thus far on this unique peroxidase have primarily focused on its unusual substrate preference for bromide (Br(-)) and the pseudohalide thiocyanate (SCN(-)) forming potent hypohalous acids as cytotoxic oxidants. However, the ability of EPO to generate reactive nitrogen species has not yet been reported. We now demonstrate that EPO readily uses nitrite (NO(2)(-)), a major end-product of nitric oxide ((.)NO) metabolism, as substrate to generate a reactive intermediate that nitrates protein tyrosyl residues in high yield. EPO-catalyzed nitration of tyrosine occurred more readily than bromination at neutral pH, plasma levels of halides, and pathophysiologically relevant concentrations of NO(2)(-). Furthermore, EPO was significantly more effective than MPO at promoting tyrosine nitration in the presence of plasma levels of halides. Whereas recent studies suggest that MPO can also promote protein nitration through indirect oxidation of NO(2)(-) with HOCl, we found no evidence that EPO can indirectly mediate protein nitration by a similar reaction between HOBr and NO(2)(-). EPO-dependent nitration of tyrosine was modulated over a physiologically relevant range of SCN(-) concentrations and was accompanied by formation of tyrosyl radical addition products (e.g. o,o'-dityrosine, pulcherosine, trityrosine). The potential role of specific antioxidants and nucleophilic scavengers on yields of tyrosine nitration and bromination by EPO are examined. Thus, EPO may contribute to nitrotyrosine formation in inflammatory conditions characterized by recruitment and activation of eosinophils.  相似文献   

8.
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   

9.
Abu-Soud HM  Hazen SL 《Biochemistry》2001,40(36):10747-10755
Recent studies demonstrate that myeloperoxidase (MPO), eosinophil peroxidase (EPO), and lactoperoxidase (LPO), homologous members of the mammalian peroxidase superfamily, can all serve as catalysts for generating nitric oxide- (nitrogen monoxide, NO) derived oxidants. These enzymes contain heme prosthetic groups that are ligated through a histidine nitrogen and use H(2)O(2) as the electron acceptor in the catalysis of oxidative reactions. Here we show that heme reduction of these peroxidases results in distinct electronic and/or conformational changes in their heme pockets using a combination of rapid kinetics measurements, optical absorbance, and diatomic ligand binding studies. Addition of reducing agent to each peroxidase at ground state [Fe(III) state] causes immediate buildup of the corresponding Fe(II) complexes. Spectral changes indicate that two LPO-Fe(II) species are present in solution at equilibrium. Analyses of stopped-flow traces collected when EPO, MPO, or LPO solutions rapidly mixed with NO were accurately fit by single-exponential functions. Plots of the apparent rate constants as a function of NO concentration for all Fe(III) and Fe(II) forms were linear with positive intercepts, consistent with NO binding to each form in a simple reversible one-step mechanism. Fe(II) forms of MPO and LPO, but not EPO, displayed significantly lower affinity toward NO compared to Fe(III) forms, suggesting that heme reduction causes a dramatic change in the heme pocket electronic environment that alters the affinity and/or accessibility of heme iron toward NO. Optical absorbance spectra indicate that CO binds to the Fe(II) forms of both LPO and EPO, but not with MPO, and generates their respective low-spin six-coordinate complexes. Kinetic analyses indicate that the binding of CO to EPO is monophasic while CO binding to LPO is biphasic. Collectively, these results illustrate for the first time functional differences in the heme pocket environments of Fe(II) forms of EPO, LPO, and MPO toward binding of diatomic ligands. Our results suggest that, upon reduction, the heme pocket of MPO collapses, LPO adopts two spectroscopically and kinetically distinguishable forms (one partially open and the other relatively closed), and EPO remains open.  相似文献   

10.
11.
Eosinophil infiltration and degranulation around the tissue-invasive stages of several species of helminths have been observed. Release of eosinophil granule contents upon the worms is supported by localization of two of the major granule proteins, major basic protein (MBP) and eosinophil peroxidase (EPO), on and around species of trematodes, nematodes, and cestodes. In the case of filarial worms, MBP is deposited on degenerating microfilariae (mf) of Onchocerca volvulus. Here, we performed in vitro assays of the toxicity of four purified eosinophil granule proteins, namely, MBP, EPO, eosinophil cationic protein (ECP), and eosinophil-derived neurotoxin (EDN), for the mf of Brugia pahangi and Brugia malayi. MBP, ECP, and EDN killed these worms in a dose-related manner although relatively high concentrations of EDN were necessary. EPO, in the presence of a H2O2-generating system and a halide, was the most potent toxin on a molar basis; here, the most potent halide was I- followed by Br- and Cl-. Surprisingly, EPO in the absence of H2O2 killed mf at concentrations comparable to those required for MBP and ECP. The toxicity of EPO + H2O2 + halide was inhibited by heparin, catalase, or 1% BSA, whereas the toxicity of EPO alone was inhibited only by heparin. Heparin also inhibited killing by both MBP and ECP. Despite the homology of ECP with certain RNases, placental RNasin, an RNase inhibitor, was unable to inhibit ECP-mediated toxicity. These results indicate that all of the eosinophil granule proteins are toxic to mf and they support the hypothesis that eosinophil degranulation causes death of mf in vivo.  相似文献   

12.
Recent studies demonstrate that nitric oxide (NO) serves as a physiological substrate for mammalian peroxidases [(2000) J. Biol. Chem. 275, 37524]. We now show that eosinophil peroxidase (EPO) and lactoperoxidase (LPO), peroxidases known to be enriched in airways of asthmatic subjects, function as a catalytic sink for NO, modulating its bioavailability and function. Using NO-selective electrodes and direct spectroscopic and rapid kinetic methods, we examined the interactions of NO with EPO and LPO compounds I and II and ferric forms and compared the results to those reported for myeloperoxidase. A unified kinetic model for NO interactions with intermediates of mammalian peroxidases during steady-state catalysis is presented that accommodates unique features observed with each member of the mammalian peroxidase superfamily. Potential functional consequences of peroxidase-NO interactions in asthma are investigated by utilizing organ chamber studies with tracheal rings. In the presence of pathophysiologically relevant levels of peroxidases and H(2)O(2), NO-dependent bronchodilation of preconstricted tracheal rings was reversibly inhibited. Thus, NO interaction with mammalian peroxidases may serve as a potential mechanism for modulating their catalytic activities, influencing the regulation of local inflammatory and infectious events in vivo.  相似文献   

13.
The kinetic characterization of horseradish peroxidase (HRPC) substrates is difficult because the reaction products are free radicals. The application of a spectrophotometrical method, which is based on determining the time necessary for a given quantity of L-ascorbic acid to be consumed (lag period) during its reaction with the free radicals generated by the enzyme acting on the reducing substrate, makes it possible to obtain the initial steady-state rates (v0). From the kinetic study of a series of derivates of phenol and aniline, the following parameters were determined for the first time: the global catalytic constant (kcat), the Michaelis constant of HRPC for H2O2 in the presence of each reducing substrate (K(M)H2O2), the Michaelis constant of HRPC for the reducing substrate (KMS), the binding constant of the reducing substrate with HRPC compound II (k5) and the rate constant of substrate oxidation by HRPC compound II (k6). The values obtained are disccussed.  相似文献   

14.
A radical species of monochlorodimedone has been characterized by its high reactivity with molecular O2. Horseradish peroxidase greatly accelerated O2 uptake by acidic solutions of this substrate; the enzymatic reaction required exogenous H2O2 only with freshly prepared substrate solutions, and the total substrate oxidized was equal to the sum of H2O2 added and O2 consumed. However, with excess Br- and horseradish peroxidase, or high Br- or Cl- and chloroperoxidase, a 1:1 stoichiometry between H2O2 and substrate was observed. In the absence of halide, the stoichiometry of the chloroperoxidase-catalyzed oxidation of monochlorodimedone changed to two molecules of the organic donor per H2O2. Moreover, in the absence of halide, at substrate:H2O2 ratios greater than 2.0, chloroperoxidase catalyzed significant O2 uptake; this enzyme-dependent autoxidation of monochlorodimedone also occurred in the presence of Cl- or Br-, when H2O2 was limiting. These data, and recent evidence from this laboratory for free hypohalous acid as the first product of chloroperoxidase-catalyzed halide oxidation [B. W. Griffin (1983) Biochem. Biophys. Res. Commun. 116, 873-879], strongly support a mixed enzymatic/nonenzymatic radical chain process as the mechanism for halogenation of monochlorodimedone by chloroperoxidase. Both horseradish peroxidase and chloroperoxidase can catalyze either bromination or oxidation of this substrate, depending on the experimental conditions. Implications of these results for the mechanism of HOCl formation catalyzed by chloroperoxidase are considered.  相似文献   

15.
Friedreich's ataxia is associated with a deficiency in frataxin, a conserved mitochondrial protein of unknown function. Here, we investigate the iron binding and oxidation chemistry of Escherichia coli frataxin (CyaY), a homologue of human frataxin, with the aim of better understanding the functional properties of this protein. Anaerobic isothermal titration calorimetry (ITC) demonstrates that at least two ferrous ions bind specifically but relatively weakly per CyaY monomer (K(d) approximately 4 microM). Such weak binding is consistent with the hypothesis that the protein functions as an iron chaperone. The bound Fe(II) is oxidized slowly by O(2). However, oxidation occurs rapidly and completely with H(2)O(2) through a non-enzymatic process with a stoichiometry of two Fe(II)/H(2)O(2), indicating complete reduction of H(2)O(2) to H(2)O. In accord with this stoichiometry, electron paramagnetic resonance (EPR) spin trapping experiments indicate that iron catalyzed production of hydroxyl radical from Fenton chemistry is greatly attenuated in the presence of CyaY. The Fe(III) produced from oxidation of Fe(II) by H(2)O(2) binds to the protein with a stoichiometry of six Fe(III)/CyaY monomer as independently measured by kinetic, UV-visible, fluorescence, iron analysis and pH-stat titrations. However, as many as 25-26 Fe(III)/monomer can bind to the protein, exhibiting UV absorption properties similar to those of hydrolyzed polynuclear Fe(III) species. Analytical ultracentrifugation measurements indicate that a tetramer is formed when Fe(II) is added anaerobically to the protein; multiple protein aggregates are formed upon oxidation of the bound Fe(II). The observed iron oxidation and binding properties of frataxin CyaY may afford the mitochondria protection against iron-induced oxidative damage.  相似文献   

16.
Single turnover reactions of the inducible nitric oxide synthase oxygenase domain (iNOSoxy) in the presence of several non alpha-amino acid N-hydroxyguanidines and guanidines were studied by stopped-flow visible spectroscopy, and compared with reactions using the native substrates L-arginine (L-arg) or N(omega)-hydroxy-L-arginine (NOHA). In experiments containing dihydrobiopterin, a catalytically incompetent pterin, and each of the studied substrates, L-arg, butylguanidine (BuGua), para-fluorophenylguanidine (FPhGua), NOHA, N-butyl- and N-(para-fluorophenyl)-N'-hydroxyguanidines (BuNOHG and FPhNOHG), the formation of a iron(II) heme-dioxygen intermediate (Fe(II)O2) was always observed. The Fe(II)O2 species then decayed to iron(III) iNOSoxy at rates that were dependent on the nature of the substrate. Identical reactions containing the catalytically competent cofactor tetrahydrobiopterin (BH4), iNOSoxy and the three N-hydroxyguanidines, all exhibited an initial formation of an Fe(II)O2 species that was successively converted to an Fe(III)NO complex and eventually to high-spin iron(III) iNOSoxy. The formation and decay kinetics of the Fe(III)NO complex did not vary greatly as a function of the N-hydroxyguanidine structure, but the formation of Fe(III)NO was substoichiometric in the cases of BuNOHG and FPhNOHG. Reactions between BH4-containing iNOSoxy and BuGua exhibited kinetics similar to those of the corresponding reaction with L-arginine, with formation of an Fe(II)O2 intermediate that was directly converted to high-spin iron(III) iNOSoxy. In contrast, no Fe(II)O2 intermediate was observed in the reaction of BH4-containing iNOSoxy and FPhGua. Multi-turnover reaction of iNOS with FPhGua did not lead to formation of NO or to hydroxylation of the substrate, contrary to reactions with BuGua or L-arg. Our results reveal how different structural and chemical properties of NOS substrate analogues can impact on the kinetics and reactivity of the Fe(II)O2 intermediate, and support an important role for substrate pKa during NOS oxygen activation.  相似文献   

17.
Eosinophil peroxidase (EPO) and lactoperoxidase (LPO) are important constituents of the innate immune system of mammals. These heme enzymes belong to the peroxidase-cyclooxygenase superfamily and catalyze the oxidation of thiocyanate, bromide and nitrite to hypothiocyanate, hypobromous acid and nitrogen dioxide that are toxic for invading pathogens. In order to gain a better understanding of the observed differences in substrate specificity and oxidation capacity in relation to heme and protein structure, a comprehensive spectro-electrochemical investigation was performed. The reduction potential (E°′) of the Fe(III)/Fe(II) couple of EPO and LPO was determined to be −126 mV and −176 mV, respectively (25 °C, pH 7.0). Variable temperature experiments show that EPO and LPO feature different reduction thermodynamics. In particular, reduction of ferric EPO is enthalpically and entropically disfavored, whereas in LPO the entropic term, which selectively stabilizes the oxidized form, prevails on the enthalpic term that favors reduction of Fe(III). The data are discussed with respect to the architecture of the heme cavity and the substrate channel. Comparison with published data for myeloperoxidase demonstrates the effect of heme to protein linkages and heme distortion on the redox chemistry of mammalian peroxidases and in consequence on the enzymatic properties of these physiologically important oxidoreductases.  相似文献   

18.
Nitrosylhemoglobin (HbFe(II)NO) has been detected in vivo, and its role in NO transport and preservation has been discussed. To gain insight into the potential role of HbFe(II)NO, we performed in vitro experiments to determine the effect of oxygenated red blood cells (RBCs) on the dissociation of cell-free HbFe(II)NO, using carboxyhemoglobin (HbFe(II)CO) as a comparison. Results show that the apparent half-life of the cell-free HbFe(II)CO was reduced significantly in the presence of RBCs at 1% hematocrit. In contrast, RBC did not change the apparent half-life of extracellular HbFe(II)NO, but caused a shift in the HbFe(II)NO dissociation product from methemoglobin (metHbFe(III)) to oxyhemoglobin (HbFe(II)O(2)). Extracellular hemoglobin was able to extract CO from HbFe(II)CO-containing RBC, but not NO from HbFe(II)NO-containing RBC. Although these results appear to suggest some unusual interactions between HbFe(II)NO and RBC, the data are explainable by simple HbFe(II)NO dissociation and hemoglobin oxidation with known rate constants. A kinetic model consisting of these reactions shows that (i) deoxyhemoglobin is an intermediate in the reaction of HbFe(II)NO oxidation to metHbFe(III), (ii) the rate-limiting step of HbFe(II)NO decay is the dissociation of NO from HbFe(II)NO, (iii) the magnitude of NO diffusion rate constant into RBC is estimated to be approximately 10(4)M(-1)s(-1), consistent with previous results determined from a competition assay, and (iv) no additional chemical reactions are required to explain these data.  相似文献   

19.
We recently described that horseradish peroxidase (HRP) and myeloperoxidase (MPO) catalyze the oxidation of melatonin, forming the respective indole ring-opening product N(1)-acetyl-N(2)-formyl-5-methoxykynuramine (AFMK) (Biochem. Biophys. Res. Commun. 279, 657-662, 2001). Although the classic peroxidatic enzyme cycle is expected to participate in the oxidation of melatonin, the requirement of a low HRP:H(2)O(2) ratio suggested that other enzyme paths might also be operative. Here we followed the formation of AFMK under two experimental conditions: predominance of HRP compounds I and II or presence of compound III. Although the consumption of substrate is comparable under both conditions, AFMK is formed in significant amounts only when compound III predominates during the reaction. Using tryptophan as substrate, N- formyl-kynurenine is formed in the presence of compound III. Both, melatonin and tryptophan efficiently prevents the formation of p-670, the inactive form of HRP. Since superoxide dismutase (SOD) inhibits the production of AFMK, we proposed that compound III acts as a source of O(-*)(2) or participates directly in the reaction, as in the case of enzyme indoleamine 2,3-dioxygenase.  相似文献   

20.
Myeloperoxidase (MPO) binds H2O2 in the absence and presence of chloride (Cl-) and catalyzes the formation of potent oxidants through 1e(-) and 2e(-) oxidation pathways. These potent oxidants have been implicated in the pathogenesis of various diseases including atherosclerosis, asthma, arthritis, and cancer. Thus, inhibition of MPO and its by-products may have a wide application in biological systems. Using direct rapid kinetic measurements and H2O2-selective electrodes, we show that tryptophan (Trp), an essential amino acid, is linked kinetically to the inhibition of MPO catalysis under physiological conditions. Trp inactivated MPO in the absence and presence of plasma levels of Cl(-), to various degrees, through binding to MPO, forming the inactive complexes Trp-MPO and Trp-MPO-Cl, and accelerating formation of MPO Compound II, an inactive form of MPO. Inactivation of MPO was mirrored by the direct conversion of MPO-Fe(III) to MPO Compound II without any sign of Compound I accumulation. This behavior indicates that Trp binding modulates the formation of MPO intermediates and their decay rates. Importantly, Trp is a poor substrate for MPO Compound II and has no role in destabilizing complex formation. Thus, the overall MPO catalytic activity will be limited by: (1) the dissociation of Trp from Trp-MPO and Trp-MPO-Cl complexes, (2) the affinity of MPO Compound I toward Cl(-) versus Trp, and (3) the slow conversion of MPO Compound II to MPO-Fe(III). Importantly, Trp-dependent inhibition of MPO occurred at a wide range of concentrations that span various physiological and supplemental ranges.  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