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.
Friedman J  Lad L  Li H  Wilks A  Poulos TL 《Biochemistry》2004,43(18):5239-5245
The Gram-negative bacterium Pseudomonas aeruginosa contains a heme oxygenase (pa-HO) that primarily oxygenates the delta-meso heme carbon [Caignan, G. A., Deshmukh, R., Wilks, A., Zeng, Y., Huang, H. W., Moenne-Loccoz, P., Bunce, R. A., Eastman, M. A., and Rivera, M. (2002) J. Am. Chem. Soc. 124, 14879-14892]. This differs from other previously characterized heme oxygenases, which display regioselectivity for the alpha-meso heme carbon. Here we report the crystal structure of pa-HO at 1.60 A resolution and compare it to the 1.50 A structure of nm-HO from Neisseria meningitidis [Schuller, D. J., Zhu, W., Stojiljkovic, I., Wilks, A., and Poulos, T. L. (2001) Biochemistry 40, 11552-11558]. The crystal structure of pa-HO maintains the same overall fold as other bacterial and mammalian heme oxygenases, including a conserved network of hydrogen-bonded solvent molecules important for dioxygen activation. The novel delta-regioselectivity of heme oxygenation observed by pa-HO is due to the heme being rotated by approximately 100 degrees, which places the delta-meso heme carbon in the same position as the alpha-meso heme carbon in other heme oxygenases. The main interaction in pa-HO that stabilizes the unique heme orientation is a salt bridge between Lys132 and the heme 7-propionate, as well as hydrophobic contacts involving Leu29, Val33, and Phe189 with the heme methyl and vinyl groups.  相似文献   

2.
Heme oxygenases from the bacterial pathogens Neisseriae meningitidis (nm-HO) and Pseudomonas aeruginosa (pa-HO) share significant sequence identity (37%). In nm-HO, biliverdin IXalpha is the sole product of the reaction, whereas pa-HO yields predominantly biliverdin IXdelta. We have previously shown by NMR that the in-plane conformation of the heme in pa-HO is significantly different from that of nm-HO as a result of distinct interactions of the heme propionates with the protein scaffold [Caignan, G. A., Deshmukh, R., Wilks, A., Zeng, Y., Huang, H. W., Moenne-Loccoz, P., Bunce, R. A., Eastman, M. A., and Rivera, M. (2002) J. Am. Chem. Soc. 124, 14879-14892]. In the report presented here, we have extended these studies to investigate the role of the distal helix by preparing a chimera of nm-HO (nm-HOch), in which distal helix residues 107-142 of nm-HO have been replaced with the corresponding residues of the delta-regioselective pa-HO (112-147). Electronic absorption spectra, resonance Raman and FTIR spectroscopic studies confirm that the orientation and hydrogen bonding properties of the proximal His ligand are not significantly altered in the chimera relative those of the wild-type proteins. The catalytic turnover of the nm-HOch-heme complex yields almost exclusively alpha-biliverdin and a small but reproducible amount of delta-biliverdin. NMR spectroscopic studies reveal that the altered regioselectivity in the chimeric protein likely stems from a dynamic equilibrium between two alternate in-plane conformations of the heme (in-plane heme disorder). Replacement of K16 with Ala and Met31 with Lys in the chimeric protein in an effort to tune key polypeptide-heme propionate contacts largely stabilizes the in-plane conformer conducive to delta-meso hydroxylation.  相似文献   

3.
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   

4.
In the oxygenase domain of mouse inducible nitric-oxide synthase (iNOSoxy), a conserved tryptophan residue, Trp-457, regulates the kinetics and extent of l-Arg oxidation to N(omega)-hydroxy-l-arginine (NOHA) by controlling electron transfer between bound (6R)-tetrahydrobiopterin (H(4)B) cofactor and the enzyme heme Fe(II)O(2) intermediate (Wang, Z. Q., Wei, C. C., Ghosh, S., Meade, A. L., Hemann, C., Hille, R., and Stuehr, D. J. (2001) Biochemistry 40, 12819-12825). To investigate whether NOHA oxidation to citrulline and nitric oxide (NO) is regulated by a similar mechanism, we performed single turnover reactions with wild type iNOSoxy and mutants W457F and W457A. Ferrous proteins containing NOHA plus H(4)B or NOHA plus 7,8-dihydrobiopterin (H(2)B), were mixed with O(2)-containing buffer, and then heme spectral transitions and product formation were followed versus time. All three proteins formed a Fe(II)O(2) intermediate with identical spectral characteristics. In wild type, H(4)B increased the disappearance rate of the Fe(II)O(2) intermediate relative to H(2)B, and its disappearance was coupled to the formation of a Fe(III)NO immediate product prior to formation of ferric enzyme. In W457F and W457A, the disappearance rate of the Fe(II)O(2) intermediate was slower than in wild type and took place without detectable build-up of the heme Fe(III)NO immediate product. Rates of Fe(II)O(2) disappearance correlated with rates of citrulline formation in all three proteins, and reactions containing H(4)B formed 1.0, 0.54, and 0.38 citrulline/heme in wild type, W457F, and W457A iNOSoxy, respectively. Thus, Trp-457 modulates the kinetics of NOHA oxidation by iNOSoxy, and this is important for determining the extent of citrulline and NO formation. Our findings support a redox role for H(4)B during NOHA oxidation to NO by iNOSoxy.  相似文献   

5.
A full-length heme oxygenase gene from the gram-negative pathogen Neisseria meningitidis was cloned and expressed in Escherichia coli. Expression of the enzyme yielded soluble catalytically active protein and caused accumulation of biliverdin within the E. coli cells. The purified HemO forms a 1:1 complex with heme and has a heme protein spectrum similar to that previously reported for the purified heme oxygenase (HmuO) from the gram-positive pathogen Corynebacterium diphtheriae and for eukaryotic heme oxygenases. The overall sequence identity between HemO and these heme oxygenases is, however, low. In the presence of ascorbate or the human NADPH cytochrome P450 reductase system, the heme-HemO complex is converted to ferric-biliverdin IXalpha and carbon monoxide as the final products. Homologs of the hemO gene were identified and characterized in six commensal Neisseria isolates, Neisseria lactamica, Neisseria subflava, Neisseria flava, Neisseria polysacchareae, Neisseria kochii, and Neisseria cinerea. All HemO orthologs shared between 95 and 98% identity in amino acid sequences with functionally important residues being completely conserved. This is the first heme oxygenase identified in a gram-negative pathogen. The identification of HemO as a heme oxygenase provides further evidence that oxidative cleavage of the heme is the mechanism by which some bacteria acquire iron for further use.  相似文献   

6.
Gohya T  Zhang X  Yoshida T  Migita CT 《The FEBS journal》2006,273(23):5384-5399
Heme oxygenase converts heme into biliverdin, CO, and free iron. In plants, as well as in cyanobacteria, heme oxygenase plays a particular role in the biosynthesis of photoreceptive pigments, such as phytochromobilins and phycobilins, supplying biliverdin IX(alpha) as a direct synthetic resource. In this study, a higher plant heme oxygenase, GmHO-1, of Glycine max (soybean), was prepared to evaluate the molecular features of its heme complex, the enzymatic activity, and the mechanism of heme conversion. The similarity in the amino acid sequence between GmHO-1 and heme oxygenases from other biological species is low, and GmHO-1 binds heme with 1 : 1 stoichiometry at His30; this position does not correspond to the proximal histidine of other heme oxygenases in their sequence alignments. The heme bound to GmHO-1, in the ferric high-spin state, exhibits an acid-base transition and is converted to biliverdin IX(alpha) in the presence of NADPH/ferredoxin reductase/ferredoxin, or ascorbate. During the heme conversion, an intermediate with an absorption maximum different from that of typical verdoheme-heme oxygenase or CO-verdoheme-heme oxygenase complexes was observed and was extracted as a bis-imidazole complex; it was identified as verdoheme. A myoglobin mutant, H64L, with high CO affinity trapped CO produced during the heme degradation. Thus, the mechanism of heme degradation by GmHO-1 appears to be similar to that of known heme oxygenases, despite the low sequence homology. The heme conversion by GmHO-1 is as fast as that by SynHO-1 in the presence of NADPH/ferredoxin reductase/ferredoxin, thereby suggesting that the latter is the physiologic electron-donating system.  相似文献   

7.
The stress protein heme oxygenase-1 (HO-1) is induced in endothelial cells exposed to nitric oxide (NO)-releasing agents, and this process is finely modulated by thiols (Foresti, R., Clark, J. E., Green, C. J., and Motterlini R. (1997) J. Biol. Chem. 272, 18411-18417). Here, we report that up-regulation of HO-1 in aortic endothelial cells by severe hypoxic conditions (pO(2) 相似文献   

8.
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   

9.
Site-directed mutagenesis studies have shown that Asp140 in both human and rat heme oxygenase-1 is critical for enzyme activity. Here, we report the D140A mutant crystal structure in the Fe(III) and Fe(II) redox states as well as the Fe(II)-NO complex as a model for the Fe(II)-oxy complex. These structures are compared to the corresponding wild-type structures. The mutant and wild-type structures are very similar, except for the distal heme pocket solvent structure. In the Fe(III) D140A mutant one water molecule takes the place of the missing Asp140 carboxylate side-chain and a second water molecule, novel to the mutant, binds in the distal pocket. Upon reduction to the Fe(II) state, the distal helix running along one face of the heme moves closer to the heme in both the wild-type and mutant structures thus tightening the active site. NO binds to both the wild-type and mutant in a bent conformation that orients the NO O atom toward the alpha-meso heme carbon atom. A network of water molecules provides a H-bonded network to the NO ligand, suggesting a possible proton shuttle pathway required to activate dioxygen for catalysis. In the wild-type structure, Asp140 exhibits two conformations, suggesting a dynamic role for Asp140 in shuttling protons from bulk solvent via the water network to the iron-linked oxy complex. On the basis of these structures, we consider why the D140A mutant is inactive as a heme oxygenase but active as a peroxidase.  相似文献   

10.
Modification of heme·heme oxygenase by iron(III) and cobalt(II) tetrasulfonated phthalocyanines has been performed. New compounds have been isolated and their properties have been investigated by difference spectroscopy, electrophoresis, molecular weight estimation, electron paramagnetic resonance (EPR) and carboxymethylation at histidyl groups. Spectrophotometric titration data indicate the ratio of the reagents in this process to be 1:1. The visible absorption spectra show the main peak at 650 nm for the iron compound and 682 nm for the cobalt one. Electrophoresis and molecular weight estimation show both complexes to be monomers. Cobalt(II) tetrasulfonated phthalocyanine, under aerobic conditions with heme oxygenase protein, undergoes autooxidation to the cobalt(III) complex, as has been proved by EPR and spectroscopic data. Iron and cobalt phthalocyanine modified heme·heme oxygenase with excess dithionite is reduced at the phthalocyanine ligand. In the presence of oxygen, the reduction product transforms into oxygenated Fe(III)Lheme oxygenase or Co(III)heme oxygenase, respectively. Reduction of the iron(III) model complex with ascorbic acid under anaerobic conditions leads to degradation of the phthalocyanine moiety, while Co(III)heme oxygenase with ascorbic acid is reduced to Co(II)Lheme oxygenase. As has been shown by carboxymethylation of the heme oxygenase protein at the histidine residues, the predominant binding site of both phthalocyanine complexes is the heme-binding histidyl residue. There is evidence that there is a second binding site with lower affinity towards Co(II)L on the heme oxygenase protein. Iron and cobalt tetrasulfonated phthalocyanines are not able to displace heme from the heme·heme oxygenase complex. In this reaction the iron complex undergoes degradation and the cobalt one gives a hybrid compound with heme·heme oxygenaseHeme oxygenase protein complexes with iron and cobalt tetrasulfonated phthalocyanines do not exhibit activity in their oxidative degradation.  相似文献   

11.
Corynebacterium diphtheriae was examined for the ability to utilize various host compounds as iron sources. C. diphtheriae C7(-) acquired iron from heme, hemoglobin, and transferrin. A siderophore uptake mutant of strain C7 was unable to utilize transferrin but was unaffected in acquisition of iron from heme and hemoglobin, which suggests that C. diphtheriae possesses a novel mechanism for utilizing heme and hemoglobin as iron sources. Mutants of C. diphtheriae and Corynebacterium ulcerans that are defective in acquiring iron from heme and hemoglobin were isolated following chemical mutagenesis and streptonigrin enrichment. A recombinant clone, pCD293, obtained from a C7(-) genomic plasmid library complemented several of the C. ulcerans mutants and three of the C. diphtheriae mutants. The nucleotide sequence of the gene (hmuO) required for complementation was determined and shown to encode a protein with a predicted mass of 24,123 Da. Sequence analysis revealed that HmuO has 33% identity and 70% similarity with the human heme oxygenase enzyme HO-1. Heme oxygenases, which have been well characterized in eukaryotes but have not been identified in prokaryotes, are involved in the oxidation of heme and subsequent release of iron from the heme moiety. It is proposed that the HmuO protein is essential for the utilization of heme as an iron source by C. diphtheriae and that the heme oxygenase activity of HmuO is involved in the release of iron from heme. This is the first report of a bacterial gene whose product has homology to heme oxygenases.  相似文献   

12.
For many pathogenic bacteria like Pseudomonas aeruginosa heme is an essential source of iron. After uptake, the heme molecule is degraded by heme oxygenases to yield iron, carbon monoxide, and biliverdin. The heme oxygenase PigA is only induced under iron-limiting conditions and produces the unusual biliverdin isomers IXbeta and IXdelta. The gene for a second putative heme oxygenase in P. aeruginosa, bphO, occurs in an operon with the gene bphP encoding a bacterial phytochrome. Here we provide biochemical evidence that bphO encodes for a second heme oxygenase in P. aeruginosa. HPLC, (1)H, and (13)C NMR studies indicate that BphO is a "classic" heme oxygenase in that it produces biliverdin IXalpha. The data also suggest that the overall fold of BphO is likely to be the same as that reported for other alpha-hydroxylating heme oxygenases. Recombinant BphO was shown to prefer ferredoxins or ascorbate as a source of reducing equivalents in vitro and the rate-limiting step for the oxidation of heme to biliverdin is the release of product. In eukaryotes, the release of biliverdin is driven by biliverdin reductase, the subsequent enzyme in heme catabolism. Because P. aeruginosa lacks a biliverdin reductase homologue, data are presented indicating an involvement of the bacterial phytochrome BphP in biliverdin release from BphO and possibly from PigA.  相似文献   

13.
Chen Y  Panda K  Stuehr DJ 《Biochemistry》2002,41(14):4618-4625
Homodimer formation is a key step that follows heme incorporation during assembly of an active inducible nitric oxide synthase (iNOS). In cells, heme incorporation into iNOS becomes limited due to interaction between self-generated NO and cellular heme [Albakri, Q., and Stuehr, D. J. (1996) J. Biol. Chem. 271, 5414-5421]. Here we investigated if NO can regulate at points downstream in the process by inhibiting dimerization of heme-containing iNOS monomer. Heme-containing monomers were generated by treating iNOS dimer or iNOS oxygenase domain dimer (iNOSoxy) with urea. Both monomers dimerized when incubated with Arg and 6R-tetrahydrobiopterin (H4B), as shown previously [Abu-Soud, H. M., Loftus, M., and Stuehr, D. J. (1995) Biochemistry 34, 11167-11175]. The NO-releasing drug S-nitrosyl-N-acetyl-D,L-penicillamine (SNAP; 0-0.5 mM) inhibited dimerization of iNOS monomer in a dose- and time-dependent manner, without causing heme release. SNAP-pretreated monomer also did not dimerize in response to H4B plus Arg. SNAP converted Arg- and H4B-free iNOS dimer into monomer that could not redimerize, but had no effect on iNOS dimer preincubated with Arg and H4B. Anaerobic spectral analysis showed that NO from SNAP bound to the ferric heme of iNOSoxy monomer or dimer. Adding imidazole as an alternative heme ligand prevented SNAP from inhibiting iNOS monomer dimerization. We conclude that NO and related species can block iNOS dimerization at points downstream from heme incorporation. The damage to heme-containing monomer results from a reaction with the protein and appears irreversible. Although dimeric structure alone does not protect, it does enable Arg and H4B to bind and protect. Inhibition appears mediated by NO coordinating to the ferric heme iron of the monomer.  相似文献   

14.
Heme oxygenase (HO) catalyzes heme degradation by utilizing O(2) and reducing equivalents to produce biliverdin IX alpha, iron, and CO. To avoid product inhibition, the heme[bond]HO complex (heme[bond]HO) is structured to markedly increase its affinity for O(2) while suppressing its affinity for CO. We determined the crystal structures of rat ferrous heme[bond]HO and heme[bond]HO bound to CO, CN(-), and NO at 2.3, 1.8, 2.0, and 1.7 A resolution, respectively. The heme pocket of ferrous heme-HO has the same conformation as that of the previously determined ferric form, but no ligand is visible on the distal side of the ferrous heme. Fe[bond]CO and Fe[bond]CN(-) are tilted, whereas the Fe[bond]NO is bent. The structure of heme[bond]HO bound to NO is identical to that bound to N(3)(-), which is also bent as in the case of O(2). Notably, in the CO- and CN(-)-bound forms, the heme and its ligands shift toward the alpha-meso carbon, and the distal F-helix shifts in the opposite direction. These shifts allow CO or CN(-) to bind in a tilted fashion without a collision between the distal ligand and Gly139 O and cause disruption of one salt bridge between the heme and basic residue. The structural identity of the ferrous and ferric states of heme[bond]HO indicates that these shifts are not produced on reduction of heme iron. Neither such conformational changes nor a heme shift occurs on NO or N(3)(-) binding. Heme[bond]HO therefore recognizes CO and O(2) by their binding geometries. The marked reduction in the ratio of affinities of CO to O(2) for heme[bond]HO achieved by an increase in O(2) affinity [Migita, C. T., Matera, K. M., Ikeda-Saito, M., Olson, J. S., Fujii, H., Yoshimura, T., Zhou, H., and Yoshida, T. (1998) J. Biol. Chem. 273, 945-949] is explained by hydrogen bonding and polar interactions that are favorable for O(2) binding, as well as by characteristic structural changes in the CO-bound form.  相似文献   

15.
16.
Heme oxygenase is the rate limiting enzyme in heme degradation to carbon monoxide (CO), iron and bilirubin. The inducible isoform of the protein, heme oxygenase-1 (HO-1), is susceptible to up-regulation by a diverse variety of conditions and agents in mammalian tissue, leading to the common conception that HO-1 is a stress related enzyme. However, as attempts are made to unravel the mechanisms by which HO-1 is induced and as we discover that CO, iron and bilirubin may be important effector molecules, we are learning to appreciate that heme oxygenases may be central to the regulation of many physiological and pathophysiological processes besides their established function in heme catabolism. One such process may be closely linked to nitric oxide (NO). It has been demonstrated that NO and NO donors are capable of inducing HO-1 protein expression, in a mechanism depending on the de novo synthesis of RNA and protein. Thus, it is postulated that NO may serve as a signaling molecule in the modulation of the tissue stress response. This review will highlight the current ideas on the role of CO-heme oxygenase and NO-nitric oxide synthase in cell signaling and discuss how the two systems are interrelated.  相似文献   

17.
The reduction of nitrite to nitric oxide in respiratory denitrification is catalyzed by a cytochrome cd(1) nitrite reductase in Paracoccus pantotrophus (formerly known as Thiosphaera pantotropha LMD 92.63). High-resolution structures are available for the fully oxidized [Fül?p, V., Moir, J. W., Ferguson, S. J., and Hajdu, J. (1995) Cell 81, 369-377; Baker, S. C., Saunders, N. F., Willis, A. C., Ferguson, S. J., Hajdu, J., and Fül?p, V. (1997) J. Mol. Biol. 269, 440-455] and fully reduced forms of this enzyme, as well as for various intermediates in its catalytic cycle [Williams, P. A., Fül?p, V., Garman, E. F., Saunders, N. F., Ferguson, S. J., and Hajdu, J. (1997) Nature 389, 406-412]. On the basis of these structures, quantum mechanical techniques (QM), including density functional methods (DFT), were combined with simulated annealing (SA) and molecular mechanics techniques (MM) to calculate the electronic distribution of molecular orbitals in the active site during catalysis. The results show likely trajectories for electrons, protons, substrates, and products in the process of nitrite reduction, and offer an interpretation of the reaction mechanism. The calculations indicate that the redox state of the d(1) heme and charges on two histidines in the active site orchestrate catalysis locally. Binding of nitrite to the reduced iron is followed by proton transfer from His345 and His388 to one of the oxygens of nitrite, creating a water molecule and an [Fe(II)-NO(+)] complex. Valence isomerization within this complex gives [Fe(III)-NO]. The release of NO from the ferric iron is influenced by the protonation state of His345 and His388, and by the orientation of NO on the d(1) heme. Return of Tyr25 to a hydrogen-bonding position between His345 and His388 facilitates product release, but a rebinding of Tyr25 to the oxidized iron may be bypassed in steady-state catalysis.  相似文献   

18.
Heme oxygenase is the rate limiting enzyme in heme degradation to carbon monoxide (CO), iron and bilirubin. The inducible isoform of the protein, heme oxygenase-1 (HO-1), is susceptible to up-regulation by a diverse variety of conditions and agents in mammalian tissue, leading to the common conception that HO-1 is a stress related enzyme. However, as attempts are made to unravel the mechanisms by which HO-1 is induced and as we discover that CO, iron and bilirubin may be important effector molecules, we are learning to appreciate that heme oxygenases may be central to the regulation of many physiological and pathophysiological processes besides their established function in heme catabolism. One such process may be closely linked to nitric oxide (NO). It has been demonstrated that NO and NO donors are capable of inducing HO-1 protein expression, in a mechanism depending on the de novo synthesis of RNA and protein. Thus, it is postulated that NO may serve as a signaling molecule in the modulation of the tissue stress response. This review will highlight the current ideas on the role of CO-heme oxygenase and NO-nitric oxide synthase in cell signaling and discuss how the two systems are interrelated.  相似文献   

19.
Nitric-oxide synthase (NOS) catalyzes the formation of NO and citrulline from l-arginine and oxygen. However, the NO so formed has been found to auto-inhibit the enzymatic activity significantly. We hypothesized that the NO reactivity is in part controlled by hydrogen bonding between the conserved tryptophan residue (position 409 in the neuronal isoform of NOS (nNOS)) and the cysteine residue that forms the proximal bond to the heme. By using resonance Raman spectroscopy and NO as a probe of the heme environment, we show that in the W409F and W409Y mutants of the oxygenase domain of the neuronal enzyme (nNOSox), the Fe-NO bond in the Fe3+NO complex is weaker than in the wild type enzyme, consistent with the loss of a hydrogen bond on the sulfur atom of the proximal cysteine residue. The weaker Fe-NO bond in the W409F and W409Y mutants might result in a faster rate of NO dissociation from the ferric heme in the Trp-409 mutants as compared with the wild type enzyme, which could contribute to the lower accumulation of the inhibitory NO-bound complexes observed during catalysis with the Trp-409 mutants (Adak, S., Crooks, C., Wang, Q., Crane, B. R., Tainer, J. A., Getzoff, E. D., and Stuehr, D. J. (1999) J. Biol. Chem. 274, 26907-26911). The optical and resonance Raman spectra of the Fe2+NO complexes of the Trp-409 mutants differ from those of the wild type enzyme and indicate that a significant population of a five-coordinate Fe2+NO complex is present. These data show that the hydrogen bond provided by the Trp-409 residue is necessary to maintain the thiolate coordination when NO binds to the ferrous heme. Taken together our results indicate that the heme environment on the proximal side of nNOS is critical for the formation of a stable iron-cysteine bond and for the control of the electronic properties of heme-NO complexes.  相似文献   

20.
Iron (II)/2-oxoglutarate (2-OG)-dependent oxygenases catalyse oxidative reactions in a range of metabolic processes including the hydroxylation of proline and lysine residues during the post-translational modification of collagen. 2-OG oxygenases commonly require ascorbate for full activity. In the vitamin C deficient disease, scurvy, reduced activity of 2-OG oxygenases results in impaired formation of collagen. Here we report the crystal structure of bacterial proline 3-hydroxylase from Streptomyces sp., an enzyme which hydroxylates proline at position 3, the first of a 2-OG oxygenase catalysing oxidation of a free alpha-amino acid. Structures were obtained for the enzyme in the absence of iron (to 2.3A resolution, R=20.2%, Rfree=25.3%) and that complexed to iron (II) (to 2.4A resolution, R=19.8%, Rfree=22.6%). The structure contains conserved motifs present in other 2-OG oxygenases including a 'jelly roll' beta strand core and residues binding iron and 2-oxoglutarate, consistent with divergent evolution within the extended family. The structure differs significantly from many other 2-OG oxygenases in possessing a discrete C-terminal helical domain. Analysis of the structure suggests a model for proline binding and a mechanism for uncoupling of proline and 2-OG turnover.  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