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 0 毫秒
1.
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   

2.
3.
4.
Heme degradation in the heme oxygenase reaction proceeds essentially as an autocatalytic oxidation of heme which is bound to heme oxygenase; in this reaction heme acts as both the substrate and the coenzyme which activates molecular oxygen. Synthesis of heme oxygenase can be induced by heme itself, in a substrate-mediated induction.  相似文献   

5.
6.
Hmu O, a heme degradation enzyme in the pathogen Corynebacterium diphtheriae, catalyzes the oxygen-dependent conversion of hemin to biliverdin, carbon monoxide, and free iron. A bacterial expression system using a synthetic gene coding for the 215-amino acid, full-length Hmu O has been constructed. Expressed at very high levels in Escherichia coli BL21, the enzyme binds hemin stoichiometrically to form a hexacoordinate high spin hemin-Hmu O complex. When ascorbic acid is used as the electron donor, Hmu O converts hemin to biliverdin with alpha-hydroxyhemin and verdoheme as intermediates. The overall conversion rate to biliverdin is approximately 4-fold slower than that by rat heme oxygenase (HO) isoform 1. Reaction of the hemin-Hmu O complex with hydrogen peroxide yields a verdoheme species, the recovery of which is much less compared with rat HO-1. Reaction of the hemin complex with meta-chloroperbenzoic acid generates a ferryl oxo species. Thus, the catalytic intermediate species and the nature of the active form in the first oxygenation step of Hmu O appear to be similar to those of the mammalian HO. However, the considerably slow catalytic rate and low level of verdoheme recovery in the hydrogen peroxide reaction suggest that the active-site structure of Hmu O is different from that of its mammalian counterpart.  相似文献   

7.
Photo-reversal of the carbon monoxide inhibition of heme oxygenase reaction by monochromatic light was investigated. Heme degradation in either the microsomal or the reconstituted heme oxygenase system was inhibited by CO. In both systems the extents of Co inhibition were dependent on the CO/O2 ratio and were nearly equal at a given CO/O2 ratio. In the reconstituted heme oxygenase reaction using a highly purified heme oxygenase preparation the relationship between the intensity of light and the degree of reversal of the CO inhibition of heme degradation as expressed in terms of delta K/Kd was not linear, but the tentatively obtained photochemical action spectrum exhibited the peaks of reversal at about 420, 540, 570, and 640 nm and suggested the occurrence of at least two steps of CO inhibition in the overall sequence of heme degradation. One could be ascribed to protoheme and the other was supposed to be the 688 nm compound which is an intermediate locating between hydroxyheme and the biliverdin-iron complex in the sequence of heme degradation.  相似文献   

8.
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   

9.
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   

10.
Baer BR  Kunze KL  Rettie AE 《Biochemistry》2007,46(41):11598-11605
Cytochrome P450s in the CYP4 family covalently bind their heme prosthetic group to a conserved acidic I-helix residue via an autocatalytic oxidation. This study was designed to evaluate the source of oxygen atoms in the covalent ester link in CYP4B1 enzymes labeled with [18O]glutamate and [18O]aspartate. The fate of the heavy isotope was then traced into wild-type CYP4B1 or the E310D mutant-derived 5-hydroxyhemes. Glutamate-containing tryptic peptides of wild-type CYP4B1 were found labeled to a level of 11-13% 18O. Base hydrolysis of labeled protein released 5-hydroxyheme which contained 12.8 +/- 1.9% 18O. Aspartate-containing peptides of the E310D mutant were labeled with 6.0-6.5% 18O, but as expected, no label was transmitted to recovered 5-hydroxyheme. These data demonstrate that the oxygen atom in 5-hydroxyheme derived from wild-type CYP4B1 originates in Glu310. Stoichiometric incorporation of the heavy isotope from the wild-type enzyme supports a perferryl-initiated carbocation mechanism for covalent heme formation in CYP4B1.  相似文献   

11.
A sensitive and facile assay for heme oxygenase (HO) has been developed. The basis of the assay is the detection of [14C]bilirubin formation in a coupled enzyme assay involving HO and biliverdin reductase actions, respectively. Separation of substrate from product is accomplished by thin-layer chromatography with subsequent quantitation by liquid scintillation counting of radioactive material present on chromatograms. As little as 20 micrograms of total cellular protein derived from cells growing in a standard 25-cm2 culture flask is sufficient for detection of HO enzyme activity using this assay. The reaction is inhibited by tin-protoporphyrin (10 microM final concentration), a specific inhibitor of HO. The linearity of the enzyme reaction with respect to incubation time and amount of protein used was established. Comparison of the new HO assay with a spectrophotometric assay was made, and good agreement of the results from both methods was found. The assay described here should facilitate measurements of this important heme-degrading enzyme in tissue culture studies and cases where limited amounts of material are available.  相似文献   

12.
13.
On incubation of peritoneal macrophages with antibody-coated radiolabeled erythrocytes, a reproducible fraction of the erythrocytes was phagocytized and heme oxygenase was induced. Addition of cyclic AMP, dbcyclic AMP, or theophylline to the incubation medium suppressed the substrate-mediated induction of heme oxygenase in a dose-related manner but did not impair the rate or extent of erythrophagocytosis. A similar effect was produced by epinephrine, norepinephrine, isoproterenol, and prostaglandins, which generate endogenous cyclic AMP by stimulating the adenyl cyclase system. Propanolol completely blocked the suppressive effect of epinephrine, while phentolamine was ineffective. In contrast to the cyclic adenosine nucleotide, cyclic GMP probably slightly enhanced the substrate-mediated induction of heme oxygenase and partly reversed the suppressive effect of cyclic AMP. Cyclic adenosine nucleotides, prostaglandin, and theophylline significantly reduced the incorporation of labeled uridine or leucine into RNA and protein of erythrophagocytic macrophages, but failed to impair the uptake of these precursors by the phagocytizing cells. These compounds also reduced the conversion of [1-14C] glucose to 14CO2 by the incubated macrophages, whereas 14CO2 formation was enhanced by epinephrine. None of these effects was reversible by addition of insulin or by glucose supplementation, which is in sharp contrast to the suppressive effect of glucocorticoids on heme oxygenase induction.  相似文献   

14.
Nitric oxide and S-nitrosothiols modulate a variety of important physiological activities. In vascular cells, agents that release NO and donate nitrosonium cation (NO(+)), such as S-nitrosoglutathione, are potent inducers of the antioxidant protein heme oxygenase 1 (HO-1) (Foresti, R., Clark, J. E., Green, C. J., and Motterlini, R. (1997) J. Biol. Chem. 272, 18411-18417; Motterlini, R., Foresti, R., Bassi, R., Calabrese, V., Clark, J. E., and Green, C. J. (2000) J. Biol. Chem. 275, 13613-13620). Here, we report that Angeli's salt (AS) (0.25-2 mm), a compound that releases nitroxyl anion (NO(-)) at physiological pH, induces HO-1 mRNA and protein expression in a concentration- and time-dependent manner, resulting in increased heme oxygenase activity in rat H9c2 cells. A time course analysis revealed that NO(-)-mediated HO-1 expression is transient and gradually disappears within 24 h, in accordance with the short half-life of AS at 37 degrees C (t(12) = 2.3 min). Interestingly, multiple additions of AS at lower concentrations (50 or 100 microm) over a period of time still promoted a significant increase in heme oxygenase activity. Experiments performed using a NO scavenger and the NO electrode confirmed that NO(-), not NO, is the species involved in HO-1 induction by AS; however, the effect on heme oxygenase activity can be amplified by accelerating the rate of NO(-) oxidation. N-Acetylcysteine almost completely abolished AS-mediated induction of HO-1, whereas a glutathione synthesis inhibitor (buthionine sulfoximine) significantly decreased heme oxygenase activation by AS, indicating that sulfydryl groups are crucial targets in the regulation of HO-1 expression by NO(-). We conclude that NO(-), in analogy with other reactive nitrogen species, is a potent inducer of heme oxygenase activity and HO-1 protein expression. These findings indicate that heme oxygenase can act both as a sensor to and target of redox-based mechanisms involving NO and extend our knowledge on the biological function of HO-1 in response to nitrosative stress.  相似文献   

15.
O2-dependent reactions of the ferric and ferrous forms of alpha-hydroxyheme complexed with water-soluble rat heme oxygenase-1 were examined by rapid-scan stopped-flow measurements. Ferric alpha-hydroxyheme reacted with O2 to form ferric verdoheme with an O2-dependent rate constant of 4x10(5) M(-1) s(-1) at pH 7.4 and 9.0. A decrease of the rate constant to 2.8x10(5) M(-1) s(-1) at pH 6.5 indicates that the reaction proceeds by direct attack of O2 on the pi-neutral radical form of alpha-hydroxyheme, which is generated by deprotonation of the alpha-hydroxy group. The reaction of ferrous alpha-hydroxyheme with O2 yielded ferrous verdoheme in a biphasic fashion involving a new intermediate having absorption maxima at 415 and 815 nm. The rate constants for this two-step reaction were 68 and 145 s(-1). These results show that conversion of alpha-hydroxyheme to verdoheme is much faster than the reduction of coordinated iron (<1 s(-1)) under physiological conditions [Y. Liu, P.R. Ortiz de Montellano, Reaction intermediates and single turnover rate constants for the oxidation of heme by human heme oxygenase-1, J. Biol. Chem. 275 (2000) 5297-5307], suggesting that, in vivo, the conversion of ferric alpha-hydroxyheme to ferric verdoheme precedes the reduction of ferric alpha-hydroxyheme.  相似文献   

16.
Microsomal heme oxygenase. Characterization of the enzyme   总被引:60,自引:0,他引:60  
  相似文献   

17.
Induction of hepatic heme oxygenase activity by bromobenzene   总被引:2,自引:0,他引:2  
Hepatic heme oxygenase, an enzyme which converts heme to carbon monoxide and bile pigment in vitro, is inducible by heme but also by large “toxic” doses of such nonheme substances as hormones, endotoxin, and heavy metal ions. When we gave rats a single hepatotoxic dose of allyl alcohol, ethionine, acetaminophen, furosemide, or endotoxin, hepatic heme oxygenase activity rose modestly (two- to fivefold) after 20 h. In contrast, administration of bromobenzene (5 mmol/kg) induced heme oxygenase in the liver an average of 15-fold after 20 h but was without effect on the enzyme in the kidney or spleen. The change in heme oxygenase was accompanied by a loss in cytochrome P-450 concentration and, in rats labeled with 5-δ-amino[14C]levulinic acid, an increased rate of degradation of hepatic [14C]heme to 14CO. Induction of heme oxygenase by bromobenzene was blocked by cycloheximide, an inhibitor of protein synthesis, but not by actinomycin D, an inhibitor of RNA synthesis. This suggests that bromobenzene stimulates de novo enzyme synthesis at the step of translation. Subtoxic doses of bromobenzene (less than 1 mmol/kg) gave proportionately greater induction of heme oxygenase. Furthermore, induction of the enzyme remained unaffected when bromobenzene hepatotoxicity was blocked by pretreatment of rats with SKF-525A, 3-methylcholanthrene, or cysteine (which supplements liver sulfhydryl content), or when hepatotoxicity was enhanced by pretreatment with phenobarbital or with diethylmaleate (which depletes hepatic glutathione). These data suggest that with induction of heme oxygenase by bromobenzene, neither liver cell necrosis nor alteration in hepatic sulfhydryl metabolism is indispensible. The latter characteristic differs from induction of the enzyme by metal ions in which depletion of sulfhydryl-containing constituents has been thought to be essential. We conclude that bromobenzene is a novel inducer of heme oxygenase activity in the liver, differing from other nonheme substances in potency and specificity for the liver, and in utilizing mechanism(s) which require neither production of hepatotoxicity, depletion of hepatic glutathione, nor sensitivity to actinomycin D.  相似文献   

18.
Prostaglandins (PGs) originate from the degradation of membranar arachidonic acid by cyclooxygenases (COX-1 and COX-2). The prostaglandin actions in the nervous system are multiple and have been suggested to play a significant role in neurodegenerative disorders. Some PGs have been reported to be toxic and, interestingly, the cyclopentenone PGs have been reported to be cytoprotective at low concentration and could play a significant role in neuronal plasticity. They have been shown to be protective against oxidative stress injury; however, the cellular mechanisms of protection afforded by these PGs are still unclear. It is postulated that the cascade leading to neuronal cell death in acute and chronic neurodegenerative conditions, such as cerebral ischemia and Alzheimer's disease, would be mediated by free radical damage. We tested the hypothesis that the neuroprotective action of cyclopentanone could be caused partially by an induction of heme oxygenase 1 (HO-1). We and others have previously reported that modulation of HO total activity may well have direct physiological implications in stroke and in Alzheimer's disease. HO acts as an antioxidant enzyme by degrading heme into iron, carbon monoxide, and biliverdin that is rapidly converted into bilirubin. Using mouse primary neuronal cultures, we demonstrated that PGs of the J series induce HO-1 in a dose-dependent manner (0, 0.5, 5, 10, 20, and 50 micro g/ml) and that PGJ(2) and dPGJ(2) were more potent than PGA(2), dPGA(2), PGD(2), and PGE(2). No significant effects were observed for HO-2 and actin expression. In regard to HO-3 expression found in rat, with its protein deducted sequence highly homologous to HO-2, no detection was observed in HO-2(-/-) mice, suggesting that HO-3 protein would not be present in mouse brain. We are proposing that several of the protective effects of PGJ(2) could be mediated through beneficial actions of heme degradation and its metabolites. The design of new mimetics based on the cyclopentenone structure could be very useful as neuroprotective agents and be tested in animal models of stroke and Alzheimer's disease.  相似文献   

19.
20.
The oxidative cleavage of heme to release iron is a mechanism by which some bacterial pathogens can utilize heme as an iron source. The pigA gene of Pseudomonas aeruginosa is shown to encode a heme oxygenase protein, which was identified in the genome sequence by its significant homology (37%) with HemO of Neisseria meningitidis. When the gene encoding the neisserial heme oxygenase, hemO, was replaced with pigA, we demonstrated that pigA could functionally replace hemO and allow for heme utilization by neisseriae. Furthermore, when pigA was disrupted by cassette mutagenesis in P. aeruginosa, heme utilization was defective in iron-poor media supplemented with heme. This defect could be restored both by the addition of exogenous FeSO4, indicating that the mutant did not have a defect in iron metabolism, and by in trans complementation with pigA from a plasmid with an inducible promoter. The PigA protein was purified by ion-exchange chromotography. The UV-visible spectrum of PigA reconstituted with heme showed characteristics previously reported for other bacterial and mammalian heme oxygenases. The heme-PigA complex could be converted to ferric biliverdin in the presence of ascorbate, demonstrating the need for an exogenous reductant. Acidification and high-performance liquid chromatography analysis of the ascorbate reduction products identified a major product of biliverdin IX-beta. This differs from the previously characterized heme oxygenases in which biliverdin IX-alpha is the typical product. We conclude that PigA is a heme oxygenase and may represent a class of these enzymes with novel regiospecificity.  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