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.
Nagababu E  Rifkind JM 《Biochemistry》2000,39(40):12503-12511
The reaction of Fe(II) hemoglobin (Hb) but not Fe(III) hemoglobin (metHb) with hydrogen peroxide results in degradation of the heme moiety. The observation that heme degradation was inhibited by compounds, which react with ferrylHb such as sodium sulfide, and peroxidase substrates (ABTS and o-dianisidine), demonstrates that ferrylHb formation is required for heme degradation. A reaction involving hydrogen peroxide and ferrylHb was demonstrated by the finding that heme degradation was inihibited by the addition of catalase which removed hydrogen peroxide even after the maximal level of ferrylHb was reached. The reaction of hydrogen peroxide with ferrylHb to produce heme degradation products was shown by electron paramagnetic resonance to involve the one-electron oxidation of hydrogen peroxide to the oxygen free radical, superoxide. The inhibition by sodium sulfide of both superoxide production and the formation of fluorescent heme degradation products links superoxide production with heme degradation. The inability to produce heme degradation products by the reaction of metHb with hydrogen peroxide was explained by the fact that hydrogen peroxide reacting with oxoferrylHb undergoes a two-electron oxidation, producing oxygen instead of superoxide. This reaction does not produce heme degradation, but is responsible for the catalytic removal of hydrogen peroxide. The rapid consumption of hydrogen peroxide as a result of the metHb formed as an intermediate during the reaction of reduced hemoglobin with hydrogen peroxide was shown to limit the extent of heme degradation.  相似文献   

2.
7,8-Dihydroneopterin and neopterin are secreted by human and primate macrophages after activation by interferon-gamma in a ratio of 2:1. 7,8-Dihydroneopterin is known to suppress radical-mediated processes, but it is also able in the presence of iron ions to generate superoxide radical anion and hydroxyl radicals from molecular oxygen. Effects of 7,8-dihydroneopterin were investigated on (met)myoglobin and (met)hemoglobin. Addition of 7,8-dihydroneopterin to heme proteins in air-saturated solution resulted in dose-dependent cleavage of the porphyrin moiety. The liberation of non-heme iron and carbon monoxide originating from the cleaved porphyrin was quantified. Both were generated at equimolar concentrations with a linear correlation coefficient of 0.9. Addition of ferrous iron significantly accelerated the pteridine-mediated cleaving of the porphyrin. However, the total yield of porphyrin cleaved was controlled by the pterin rather than by the ferrous ion concentration. 7,8-Dihydroneopterin is assumed to reduce the heme iron in intact protein molecules, thereby preparing the conditions for binding of oxygen and carbon monoxide as ligands. Beyond that, it is concluded that hydroxyl radicals might be generated via reduction of molecular oxygen to superoxide anion in the autoxidation process and dismutation to hydrogen peroxide and subsequent Fenton reaction.  相似文献   

3.
Hypochlorous acid (HOCl) is generated by myeloperoxidase using chloride and hydrogen peroxide as substrates. HOCl and its conjugate base (OCl) bind to the heme moiety of hemoglobin (Hb) and generate a transient ferric species whose formation and decay kinetics indicate it can participate in protein aggregation and heme destruction along with subsequent free iron release. The oxidation of the Hb heme moiety by OCl was accompanied by marked heme destruction as judged by the decrease in and subsequent flattening of the Soret absorbance peak at 405 nm. HOCl-mediated Hb heme depletion was confirmed by HPLC analysis and in-gel heme staining. Exposure of Hb to increasing concentrations of HOCl produced a number of porphyrin degradation products resulting from oxidative cleavage of one or more of the carbon-methene bridges of the tetrapyrrole ring, as identified by their characteristic HPLC fluorescence and LC-MS. A nonreducing denaturing SDS-PAGE showed several degrees of protein aggregation. Similarly, porphyrin degradation products were identified after exposure of red blood cells to increasing concentrations of HOCl, indicating biological relevance of this finding. This work provides a direct link between Hb heme destruction and subsequent free iron accumulation, as occurs under inflammatory conditions where HOCl is formed in substantial amounts.  相似文献   

4.
Glucose modifies the amino groups of proteins by a process of non-enzymatic glycation, leading to potentially deleterious effects on structure and function that have been implicated in the pathogenesis of diabetic complications. These changes are extremely complex and occur very slowly. We demonstrate here that hemoglobin and myoglobin are extremely susceptible to damage by glucose in vitro through a process that leads to complete destruction of the essential heme group. This process appears in addition to the expected formation of so-called advanced glycation end products (AGEs) on lysine and other side-chains. AGE formation is enhanced by the iron released. In contrast, the heme group is not destroyed during glycation of cytochrome c, where the sixth coordination position of the heme iron is not accessible to solvent ligands. Glycation leads to reduction of ferricytochrome c in this case. Since hydrogen peroxide is known to destroy heme, and the destruction observed during glycation of hemoglobin and myoglobin is sensitive to catalase, we propose that the degradation process is initiated by hydrogen peroxide formation. Damage may then occur through reaction with superoxide generated (a reductant of ferricytochrome c), or hydroxyl radicals, or with both.  相似文献   

5.
Physiological heme degradation is mediated by the heme oxygenase system consisting of heme oxygenase and NADPH-cytochrome P-450 reductase. Biliverdin IX alpha is formed by elimination of one methene bridge carbon atom as CO. Purified NADPH-cytochrome P-450 reductase alone will also degrade heme but biliverdin is a minor product (15%). The enzymatic mechanisms of heme degradation in the presence and absence of heme oxygenase were compared by analyzing the recovery of 14CO from the degradation of [14C]heme. 14CO recovery from purified NADPH-cytochrome P-450 reductase-catalyzed degradation of [14C]methemalbumin was 15% of the predicted value for one molecule of CO liberated per mole of heme degraded. 14CO2 and [14C]formic acid were formed in amounts (18 and 98%, respectively), suggesting oxidative cleavage of more than one methene bridge per heme degraded, similar to heme degradation by hydrogen peroxide. The reaction was strongly inhibited by catalase, but superoxide dismutase had no effect. [14C]Heme degradation by the reconstituted heme oxygenase system yielded 33% 14CO. Near-stoichiometric recovery of 14CO was achieved after addition of catalase to eliminate side reactions. Near-quantitative recovery of 14CO was also achieved using spleen microsomal preparations. Heme degradation by purified NADPH-cytochrome P-450 reductase appeared to be mediated by hydrogen peroxide. The major products were not bile pigments, and only small amounts of CO were formed. The presence of heme oxygenase, and possibly an intact membrane structure, were essential for efficient heme degradation to bile pigments, possibly by protecting the heme from indiscriminate attack by active oxygen species.  相似文献   

6.
The role of the proximal heme iron ligand in activation of hydrogen peroxide and control of spin state and coordination number in heme proteins is not yet well understood. Although there are several examples of amino acid sidechains with oxygen atoms which can act as potential heme iron ligands, the occurrence of protein-derived oxygen donor ligation in natural protein systems is quite rare. The sperm whale myoglobin cavity mutant H93G Mb (D. Barrick, Biochemistry 33 (1994) 6546) has its proximal histidine ligand replaced by glycine, a mutation which leaves an open cavity capable of accommodation of a variety of unnatural potential proximal ligands. This provides a convenient system for studying ligand-protein interactions. Molecular modeling of the proximal cavity in the active site of H93G Mb indicates that the cavity is of sufficient size to accommodate benzoate and phenolate in conformations that allow their oxygen atoms to come within binding distance of the heme iron. In addition, benzoate may occupy the cavity in an orientation which allows one carboxylate oxygen atom to ligate to the heme iron while the other carboxylate oxygen is within hydrogen bonding distance of serine 92. The ferric phenolate and benzoate complexes have been prepared and characterized by UV-visible and MCD spectroscopies. The benzoate adduct shows characteristics of a six-coordinate high-spin complex. To our knowledge, this is the first known example of a six-coordinate high-spin heme complex with an anionic oxygen donor proximal ligand. The benzoate ligand is displaced at alkaline pH and upon reaction with hydrogen peroxide. The phenolate adduct of H93G Mb is a five-coordinate high-spin complex whose UV-visible and MCD spectra are distinct from those of the histidine 93 to tyrosine (H93Y Mb) mutant of sperm whale myoglobin. The phenolate adduct is stable at alkaline pH and exhibits a reduced reactivity with hydrogen peroxide relative to that of both native ferric myoglobin, and the exogenous ligand-free derivative of ferric H93G Mb. These observations indicate that the identity of the proximal oxygen donor ligand has an important influence on both the heme iron coordination number and the reactivity of the complex with hydrogen peroxide.  相似文献   

7.
Heme oxygenase (HO) converts hemin to biliverdin, CO, and iron applying molecular oxygen and electrons. During successive HO reactions, two intermediates, α-hydroxyhemin and verdoheme, have been generated. Here, oxidation state of the verdoheme-HO complexes is controversial. To clarify this, the heme conversion by soybean and rat HO isoform-1 (GmHO-1 and rHO-1, respectively) was compared both under physiological conditions, with oxygen and NADPH coupled with ferredoxin reductase/ferredoxin for GmHO-1 or with cytochrome P450 reductase for rHO-1, and under a non-physiological condition with hydrogen peroxide. EPR measurements on the hemin-GmHO-1 reaction with oxygen detected a low-spin ferric intermediate, which was undetectable in the rHO-1 reaction, suggesting the verdoheme in the six-coordinate ferric state in GmHO-1. Optical absorption measurements on this reaction indicated that the heme degradation was extremely retarded at verdoheme though this reaction was not inhibited under high-CO concentrations, unlike the rHO-1 reaction. On the contrary, the Gm and rHO-1 reactions with hydrogen peroxide both provided ferric low-spin intermediates though their yields were different. The optical absorption spectra suggested that the ferric and ferrous verdoheme coexisted in reaction mixtures and were slowly converted to the ferric biliverdin complex. Consequently, in the physiological oxygen reactions, the verdoheme is found to be stabilized in the ferric state in GmHO-1 probably guided by protein distal residues and in the ferrous state in rHO-1, whereas in the hydrogen peroxide reactions, hydrogen peroxide or hydroxide coordination stabilizes the ferric state of verdoheme in both HOs.  相似文献   

8.
Mechanism of heme degradation by heme oxygenase   总被引:5,自引:0,他引:5  
Heme oxygenase catalyzes the three step-wise oxidation of hemin to alpha-biliverdin, via alpha-meso-hydroxyhemin, verdoheme, and ferric iron-biliverdin complex. This enzyme is a simple protein which does not have any prosthetic groups. However, heme and its two metabolites, alpha-meso-hydroxyhemin and verdoheme, combine with the enzyme and activate oxygen during the heme oxygenase reaction. In the conversion of hemin to alpha-meso-hydroxyhemin, the active species of oxygen is Fe-OOH, which self-hydroxylates heme to form alpha-meso-hydroxyhemin. This step determines the alpha-specificity of the reaction. For the formation of verdoheme and liberation of CO from alpha-meso-hydroxyhemin, oxygen and one reducing equivalent are both required. However, the ferrous iron of the alpha-meso-hydroxyheme is not involved in the oxygen activation and unactivated oxygen is reacted on the 'activated' heme edge of the porphyrin ring. For the conversion of verdoheme to the ferric iron-biliverdin complex, both oxygen and reducing agents are necessary, although the precise mechanism has not been clear. The reduction of iron is required for the release of iron from the ferric iron-biliverdin complex to complete total heme oxygenase reaction.  相似文献   

9.
In heme degradation catalyzed by the reconstituted heme oxygenase system, 8 to 9 mol of dioxygen and 11 to 12 mol of NADPH were consumed per mol of hemin lost, and about half the amount of dioxygen consumed could be accounted for by the production of hydrogen peroxide, which accumulated in the reaction mixture. Production of hydrogen peroxide in the heme oxygenase reaction did not appear to be due to the bimolecular dismutation of superoxide anions but rather seemed to be due to dissociation of a "peroxo" species formed on heme or intermediates of heme degradation. The hydrogen peroxide produced appeared to cause a considerable degree of non-specific degradation of heme (not leading to the formation of biliverdin) and also caused an inactivation of heme oxygenase. By taking into account the amount of dioxygen incorporated into hydrogen peroxide and some other factors, it could be deduced that 3 mol of dioxygen is consumed for the formation of 1 mol of biliverdin in the heme oxygenase reaction.  相似文献   

10.
Excessive release of hemoglobin from red blood cells markedly disturbs the health status of patients due to cytotoxic effects of free hemoglobin and heme. The latter component is able to initiate novel hemolytic events in unperturbed red blood cells. We modeled this process by incubation of ferric protoporphyrin IX with freshly isolated red blood cells from healthy volunteers. The heme-induced hemolysis was inhibited in a concentration-dependent manner by the chlorite-based drug WF10, whereby the hemolysis degree was totally abolished at a molar ratio of 1:2 between chlorite and heme. Upon incubation of heme with WF10, the ultraviolet-visible spectrum changed, whereas the release of iron from heme and the appearance of fluorescent breakdown products of the porphyrin ring were negligible at this ratio, but increased with increasing excess of chlorite over heme. Thus, inhibition of hemolysis by WF10 takes already place at those chlorite concentrations, where no degradation of the porphyrin ring occurs. As WF10 is applied in form of an intravenous infusion to patients with severe inflammatory states, these data support the hypothesis that the beneficial WF10 effects are closely associated with inactivation of free heme.  相似文献   

11.
Here, we show that hypochlorous acid (HOCl), a potent neutrophil-generated oxidant, can mediate destruction of free heme (Ht) and the heme precursor, protoporphyrin IX (PPIX). Ht displays a broad Soret absorbance peak centered at 365 and 394 nm, indicative of the presence of monomer and μ-oxo-dimer. Oxidation of Ht by HOCl was accompanied by a marked decrease in the Soret absorption peak and release of free iron. Kinetic measurements showed that the Ht-HOCl reaction was triphasic. The first two phases were HOCl concentration dependent and attributable to HOCl binding to the monomeric and dimeric forms. The third phase was HOCl concentration independent and attributed to Ht destruction with the release of free iron. HPLC and LC-ESI-MS analyses of the Ht-HOCl reaction revealed the formation of a number of degradation products, resulting from the cleavage or modification of one or more carbon-methene bridges of the porphyrin ring. Similar studies with PPIX showed that HOCl also mediated tetrapyrrole ring destruction. Collectively, this work demonstrates the ability of HOCl to modulate destruction of heme, through a process that occurs independent of the iron molecule that resides in the porphyrin center. This phenomenon may play a role in HOCl-mediated oxidative injury in pathological conditions.  相似文献   

12.
The copper complex of 2,9-dimethyl-1,10-phenanthroline(2,9-dmp) is accumulated by a variety of organisms and is highly toxic. Bioaccumulation depends on reduction of copper(II) to (I), since only the copper(I)-2,9-dmp complex is lipophilic. In the case of the marine diatom, Nitzschia closterium, it is proposed that hydrogen peroxide, produced by the algae during photosynthesis, is the in vivo reductant. Hydrogen peroxide rapidly reduces copper(II)-2,9-dmp, but an excess of H2O2 leads to destruction of the yellow copper(I) complex. Rate constants for the formation and degradation of the yellow complex are reported. Oxygen, light, and a hydroxylating agent are released during the degradation reaction. A reaction mechanism is proposed for the destruction of copper-2,9-dmp by excess H2O2, involving attack on the 5, 6 positions of the phenanthroline ring by hydroxyl radical, then further oxidation by singlet oxygen and H2O2. These in vivo degradation reactions are believed to be the cause of the extreme toxicity of the complex.  相似文献   

13.
14.
The formation of compound I is the first step in the reaction mechanism of plant heme peroxidases. This intermediate stores two oxidizing equivalents from hydrogen peroxide as an oxyferryl iron center and a radical, either on the porphyrin ring or on a tryptophan residue. Site-directed mutagenesis has proved to be a most useful tool for the identification of the intermediates involved and the resulting nature of the compound I formed. Although there is no doubt that an acid-base mechanism operates in heme peroxidase during the formation of compound I, the roles of several distal pocket residues are currently the subject of intensive research. It is now generally accepted that the conserved distal histidine in the active site of heme peroxidases is the acid-base catalyst that promotes the heterolytic cleavage of hydrogen peroxide. Other residues, such as the distal arginine and asparagine, participate in a range of roles assisting catalysis by the distal histidine. Recent advances in the elucidation of the mechanism at the molecular level are discussed. Another aspect related to the nature of compound I is the location of the radical center. Novel radical species have been detected in the reactions of ascorbate peroxidase, lignin peroxidase and several mutants of horseradish peroxidase. Detailed kinetic and spectroscopic studies of these radical species have provided important insights about the factors that control porphyrin-protein radical exchange. The wide range of data being obtained on compound I will lead to an understanding of its vital function in peroxidase catalysis and the physiological roles played by these enzymes.  相似文献   

15.
Protoporphyrin, an intermediate in heme and chlorophyll biosynthesis, can accumulate in human and plant tissues under certain pathological conditions and is a photosensitizer used in cancer phototherapy. We previously showed that protoporphyrin and the related non-natural dicarboxylic porphyrin deuteroporphyrin are rapidly oxidized by horseradish peroxidase in the presence of some thiols, especially glutathione. This study reports that bovine lactoperoxidase, but not leucocyte myeloperoxidase, can also catalyze this reaction and that Tween and ascorbic acid are inhibitors. Exogenous hydrogen peroxide is not required and cannot replace glutathione. Deuteroporphyrin was oxidized to a unique green chlorin product with two oxygen functions added directly to the characteristic reduced pyrrole ring of the chlorin. Spectroscopic and chromatographic results suggest that protoporphyrin was oxidized not to a green chlorin, but to a much more polar red porphyrin modified by oxidative addition to the two vinyl side chains. Two related nonnatural dicarboxylic porphyrins, with ethyl or hydroxyethyl instead of vinyl side chains, are not substrates or products for this enzymatic conversion.  相似文献   

16.
Resonance Raman spectra of native, overexpressed M. tuberculosis catalase-peroxidase (KatG), the enzyme responsible for activation of the antituberculosis antibiotic isoniazid (isonicotinic acid hydrazide), have confirmed that the heme iron in the resting (ferric) enzyme is high-spin five-coordinate. Difference Raman spectra did not reveal a change in coordination number upon binding of isoniazid to KatG. Stopped-flow spectrophotometric studies of the reaction of KatG with stoichiometric equivalents or small excesses of hydrogen peroxide revealed only the optical spectrum of the ferric enzyme with no hypervalent iron intermediates detected. Large excesses of hydrogen peroxide generated oxyferrous KatG, which was unstable and rapidly decayed to the ferric enzyme. Formation of a pseudo-stable intermediate sharing optical characteristics with the porphyrin pi-cation radical-ferryl iron species (Compound I) of horseradish peroxidase was observed upon reaction of KatG with excess 3-chloroperoxybenzoic acid, peroxyacetic acid, or tert-butylhydroperoxide (apparent second-order rate constants of 3.1 x 10(4), 1.2 x 10(4), and 25 M(-1) s(-1), respectively). Identification of the intermediate as KatG Compound I was confirmed using low-temperature electron paramagnetic resonance spectroscopy. Isoniazid, as well as ascorbate and potassium ferrocyanide, reduced KatG Compound I to the ferric enzyme without detectable formation of Compound II in stopped-flow measurements. This result differed from the reaction of horseradish peroxidase Compound I with isoniazid, during which Compound II was stably generated. These results demonstrate important mechanistic differences between a bacterial catalase-peroxidase and the homologous plant peroxidases and yeast cytochrome c peroxidase, in its reactions with peroxides as well as substrates.  相似文献   

17.
The structure-function relationship of two kinds of hemoproteins, peroxidases and peroxygenase, is discussed and a tentative model for the active site (heme vicinity) structure of each hemoprotein is proposed. The mechanism of Compound I formation from peroxidases is presumed to involve an electrophilic attack of hydroperoxide, the electrophilicity of which is increased by forming a hydrogen bond to a distal acid group (with β-equatorial arrangement) on the heme iron, the basicity of which is being increased by electron donation from the anionic fifth ligand. On the other hand, the mechanism for peroxygenase is presumed to involve a nucleophilic attack of hydroperoxide, the nucleophilicity of which is increased by forming a hydrogen bond to a distal base group (with α-axial arrangement) to the heme iron ligating the neutral fifth ligand. It is presumed that Compound I of peroxidases, which consists of porphyrin π cation radical and ferryl iron, is stabilized by a π-π type charge transfer interaction between the radical, and stacking imidazolate group (not necessarily different from the distal group) which then ionizes, and by electron donation from the anionic fifth ligand. On the other hand, Compound I of peroxygenase, which is postulated to be an oxene complex, is presumed to be stabilized by an electrostatic interaction with a strongly negative environment, and by ionization of the fifth ligand, if such can happen.  相似文献   

18.
Circular dichroism spectra of cytochrome c peroxidase from baker's yeast, those of the reduced enzyme, the carbonyl, cyanide and fluoride derivatives and the hydrogen peroxide compound, Compound I, have been recorded in the wavelength range 200 to 660 nm. All derivatives show negative Soret Cotton effects. The results suggest that the heme group is surrounded by tightly packed amino acid sidechains and that there is a histidine residue bound to the fifth coordination site of the heme iron. The native ferric enzyme is probably pentacoordinated. The circular dichroism spectra of the ligand compounds indicate that the ligands form a nonlinear bond to the heme iron as a result of steric hindrance in the vicinity of the heme. The spectrum of Compound I shows no perturbation of the porphyrin symmetry. The dichroic spectrum of the native enzyme in the far-ultraviolet wave-length region suggests that the secondary structure consists of roughly equal amounts of alpha-helical, beta-structure and unordered structure. After the removal of the heme group no great changes in the secondary structure can be observed.  相似文献   

19.
Three newly discovered non-heme bromoperoxidases isolated from marine algae were found to catalyze the production of singlet oxygen in reactions composed of the bromoperoxidase, hydrogen peroxide, and bromide. The bromoperoxidases studied were vanadium bromoperoxidase (V-BrPO) from Ascophyllum nodosum, native non-heme bromoperoxidase from Corallina vancouveriensis (which contains vanadium and iron), and the vanadium-reconstituted bromoperoxidase derivative from C. vancouveriensis. These enzyme systems generated near infrared emission, characteristic of singlet oxygen. The emission had a peak intensity near 1268 nm, was greatly increased in 2H2O-containing buffers, and was greatly decreased by the singlet oxygen quenchers, histidine and azide. The yield of singlet oxygen was approximately 80% of the theoretical yield. A unique feature of the non-heme bromoperoxidases distinct from the iron heme haloperoxidases, was the remarkable stability of the non-heme enzymes in the presence of singlet oxygen and oxidized bromine species. V-BrPO turned over multiple aliquots of 2 mM hydrogen peroxide without losing efficiency. In contrast, iron heme lactoperoxidase was completely inactivated after turnover of the first aliquot of 2 mM hydrogen peroxide, and iron heme chloroperoxidase was 50% deactivated. The profile of singlet oxygen formation by V-BrPO and the near stoichiometric yield of singlet oxygen suggest that the mechanism of singlet oxygen formation is the same as the mechanism of dioxygen formation determined by oxygen probe measurements.  相似文献   

20.
Two fluorescent heme degradation compounds are detected during autoxidation of oxyhemoglobin. These fluorescent compounds are similar to fluorescent compounds formed when hydrogen peroxide reacts with hemoglobin [E. Nagababu and J. M. Rifkind, Biochem. Biophys. Res. Commun. 247, 592-596 (1998)]. Low levels of heme degradation in the presence of superoxide and catalase are attributed to a reaction involving the superoxide produced during autoxidation. The inhibition of most of the degradation by catalase suggests that the hydrogen peroxide generated during autoxidation of oxyhemoglobin produces heme degradation by the same mechanism as the direct addition of hydrogen peroxide to hemoglobin. The formation of the fluorescent degradation products was inhibited by the peroxidase substrate, ABTS, which reduces ferrylhemoglobin to methemoglobin, indicating that ferrylhemoglobin is produced during the autoxidation of hemoglobin. It is the transient formation of this highly reactive Fe(IV) hemoglobin, which is responsible for most of the heme degradation during autoxidation.  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