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.
Human heme oxygenase-1 (hHO-1) catalyzes the NADPH-cytochrome P450 reductase-dependent oxidation of heme to biliverdin, CO, and free iron. The biliverdin is subsequently reduced to bilirubin by biliverdin reductase. Earlier kinetic studies suggested that biliverdin reductase facilitates the release of biliverdin from hHO-1 (Liu, Y., and Ortiz de Montellano, P. R. (2000) J. Biol. Chem. 275, 5297-5307). We have investigated the binding of P450 reductase and biliverdin reductase to truncated, soluble hHO-1 by fluorescence resonance energy transfer and site-specific mutagenesis. P450 reductase and biliverdin reductase bind to truncated hHO-1 with Kd = 0.4 +/- 0.1 and 0.2 +/- 0.1 microm, respectively. FRET experiments indicate that biliverdin reductase and P450 reductase compete for binding to truncated hHO-1. Mutation of surface ionic residues shows that hHO-1 residues Lys18, Lys22, Lys179, Arg183, Arg198, Glu19, Glu127, and Glu190 contribute to the binding of cytochrome P450 reductase. The mutagenesis results and a computational analysis of the protein surfaces partially define the binding site for P450 reductase. An overlapping binding site including Lys18, Lys22, Lys179, Arg183, and Arg185 is similarly defined for biliverdin reductase. These results confirm the binding of biliverdin reductase to hHO-1 and define binding sites of the two reductases.  相似文献   

2.
Heme oxygenase catalyzes the regiospecific oxidation of hemin to biliverdin IXalpha with concomitant liberation of CO and iron by three sequential monooxygenase reactions. The alpha-regioselectivity of heme oxygenase has been thought to result from the regioselective oxygenation of the heme alpha-meso position at the first step, which leads to the reaction pathway via meso-hydroxyheme IXalpha and verdoheme IXalpha intermediates. However, recent reports concerning heme oxygenase forming biliverdin isomers other than biliverdin IXalpha raise a question whether heme oxygenase can degrade meso-hydroxyhemin and isomers other than the alpha-isomers. In this paper, we investigated the stereoselectivity of each of the two reaction steps from meso-hydroxyhemin to verdoheme and verdoheme to biliverdin by using a truncated form of rat heme oxygenase-1 and the chemically synthesized four isomers of meso-hydroxyhemin and verdoheme. Heme oxygenase-1 converted all four isomers of meso-hydroxyhemin to the corresponding isomers of verdoheme. In contrast, only verdoheme IXalpha was converted to the corresponding biliverdin IXalpha. We conclude that the third step, but not the second, is stereoselective for the alpha-isomer substrate. The present findings on regioselectivities of the second and the third steps have been discussed on the basis of the oxygen activation mechanisms of these steps.  相似文献   

3.
The ability of the human heme oxygenase-1 (hHO-1) R183E mutant to oxidize heme in reactions supported by either NADPH-cytochrome P450 reductase or ascorbic acid has been compared. The NADPH-dependent reaction, like that of wild-type hHO-1, yields exclusively biliverdin IXalpha. In contrast, the R183E mutant with ascorbic acid as the reductant produces biliverdin IXalpha (79 +/- 4%), IXdelta (19 +/- 3%), and a trace of IXbeta. In the presence of superoxide dismutase and catalase, the yield of biliverdin IXdelta is decreased to 8 +/- 1% with a corresponding increase in biliverdin IXalpha. Spectroscopic analysis of the NADPH-dependent reaction shows that the R183E ferric biliverdin complex accumulates, because reduction of the iron, which is required for sequential iron and biliverdin release, is impaired. Reversal of the charge at position 183 makes reduction of the iron more difficult. The crystal structure of the R183E mutant, determined in the ferric and ferrous-NO bound forms, shows that the heme primarily adopts the same orientation as in wild-type hHO-1. The structure of the Fe(II).NO complex suggests that an altered active site hydrogen bonding network supports catalysis in the R183E mutant. Furthermore, Arg-183 contributes to the regiospecificity of the wild-type enzyme, but its contribution is not critical. The results indicate that the ascorbate-dependent reaction is subject to a lower degree of regiochemical control than the NADPH-dependent reaction. Ascorbate may be able to reduce the R183E ferric and ferrous dioxygen complexes in active site conformations that cannot be reduced by NADPH-cytochrome P450 reductase.  相似文献   

4.
Heme oxygenase regiospecifically oxidizes heme at the alpha-meso position to give biliverdin IXalpha, CO, and iron. The heme orientation within the active site, which is thought to determine the oxidation regiospecificity, is shown here for the human enzyme (hHO1) to be largely determined by interactions between the heme carboxylic acid groups and residues Arg183 and Lys18 but not Tyr134. Mutation of either Arg183 or Lys18 individually does not significantly alter the NADPH-cytochrome P450 reductase-dependent reaction regiochemistry but partially shifts the oxidation to the beta/delta-meso positions in the reaction supported by ascorbic acid. Mutation of Glu29 to a lysine, which places a positive charge where it can interact with a heme carboxyl if the heme rotates by approximately 90 degrees, causes a slight loss of regiospecificity but combined with the R183E and K18E mutations results primarily in beta/delta-meso oxidation of the heme under all conditions. NMR analysis of heme binding to the triple K18E/E29K/R183E mutant confirms rotation of the heme in the active site. Kinetic studies demonstrate that mutations of Arg183 greatly impair the rate of the P450 reductase-dependent reaction, in accord with the earlier finding that Arg183 is involved in binding of the reductase to hHO1, but have little effect on the ascorbate reaction. Mutations of Asp140 and Tyr58 that disrupt the active site hydrogen bonding network impair catalytic rates but do not influence the oxidation regiochemistry. The results indicate both that the oxidation regiochemistry is largely controlled by ionic interactions of the heme propionic acid groups with the protein and that shifts in regiospecificity involve rotation of the heme about an axis perpendicular to the heme plane.  相似文献   

5.
Human heme oxygenase-1 (hHO-1) catalyzes the O2- and NADPH-dependent oxidation of heme to biliverdin, CO, and free iron. The first step involves regiospecific insertion of an oxygen atom at the alpha-meso carbon by a ferric hydroperoxide and is predicted to proceed via an isoporphyrin pi-cation intermediate. Here we report spectroscopic detection of a transient intermediate during oxidation by hHO-1 of alpha-meso-phenylheme-IX, alpha-meso-(p-methylphenyl)-mesoheme-III, and alpha-meso-(p-trifluoromethylphenyl)-mesoheme-III. In agreement with previous experiments (Wang, J., Niemevz, F., Lad, L., Huang, L., Alvarez, D. E., Buldain, G., Poulos, T. L., and Ortiz de Montellano, P. R. (2004) J. Biol. Chem. 279, 42593-42604), only the alpha-biliverdin isomer is produced with concomitant formation of the corresponding benzoic acid. The transient intermediate observed in the NADPH-P450 reductase-catalyzed reaction accumulated when the reaction was supported by H2O2 and exhibited the absorption maxima at 435 and 930 nm characteristic of an isoporphyrin. Product analysis by reversed phase high performance liquid chromatography and liquid chromatography electrospray ionization mass spectrometry of the product generated with H2O2 identified it as an isoporphyrin that, on quenching, decayed to benzoylbiliverdin. In the presence of H218O2, one labeled oxygen atom was incorporated into these products. The hHO-1-isoporphyrin complexes were found to have half-lives of 1.7 and 2.4 h for the p-trifluoromethyl- and p-methyl-substituted phenylhemes, respectively. The addition of NADPH-P450 reductase to the H2O2-generated hHO-1-isoporphyrin complex produced alpha-biliverdin, confirming its role as a reaction intermediate. Identification of an isoporphyrin intermediate in the catalytic sequence of hHO-1, the first such intermediate observed in hemoprotein catalysis, completes our understanding of the critical first step of heme oxidation.  相似文献   

6.
Hemoglobin digestion in the midgut of hematophagous animals results in the release of its prosthetic group, heme, which is a pro-oxidant molecule. Heme enzymatic degradation is a protective mechanism that has been described in several organisms, including plants, bacteria, and mammals. This reaction is catalyzed by heme oxygenase and results in formation of carbon monoxide, ferrous ion, and biliverdin IXalpha. During digestion, a large amount of a green pigment is produced and secreted into the intestinal lumen of Aedes aegypti adult females. In the case of another blood-sucking insect, the kissing-bug Rhodnius prolixus, we have recently shown that heme degradation involves a complex pathway that generates dicysteinyl-biliverdin IX gamma. The light absorption spectrum of the Aedes purified pigment was similar to that of biliverdin, but its mobility on a reverse-phase chromatography column suggested a compound less hydrophobic than biliverdin IXalpha. Structural characterization by ESI-MS revealed that the mosquito pigment is the alpha isomer of biliverdin bound to two glutamine residues by an amide bond. This biglutaminyl-biliverdin is formed by oxidative cleavage of the heme porphyrin ring followed by two subsequent additions of glutamine residues to the biliverdin IXalpha. The role of this pathway in the adaptation of this insect vector to a blood-feeding habit is discussed.  相似文献   

7.
Interaction of nitric oxide with human heme oxygenase-1   总被引:2,自引:0,他引:2  
NO and CO may complement each other as signaling molecules in some physiological situations. We have examined the binding of NO to human heme oxygenase-1 (hHO-1), an enzyme that oxidizes heme to biliverdin, CO, and free iron, to determine whether inhibition of hHO-1 by NO can contribute to the signaling interplay of NO and CO. An Fe(3+)-NO hHO-1-heme complex is formed with NO or the NO donors NOC9 or 2-(N,N-diethylamino)-diazenolate-2-oxide.sodium salt. Resonance Raman spectroscopy shows that ferric hHO-1-heme forms a 6-coordinated, low spin complex with NO. The nu(N-O) vibration of this complex detected by Fourier transform IR is only 4 cm(-1) lower than that of the corresponding metmyoglobin (met-Mb) complex but is broader, suggesting a greater degree of ligand conformational freedom. The Fe(3+)-NO complex of hHO-1 is much more stable than that of met-Mb. Stopped-flow studies indicate that k(on) for formation of the hHO-1-heme Fe(3+)-NO complex is approximately 50-times faster, and k(off) 10 times slower, than for met-Mb, resulting in K(d) = 1.4 microm for NO. NO thus binds 500-fold more tightly to ferric hHO-1-heme than to met-Mb. The hHO-1 mutations E29A, G139A, D140A, S142A, G143A, G143F, and K179A/R183A do not significantly diminish the tight binding of NO, indicating that NO binding is not highly sensitive to mutations of residues that normally stabilize the distal water ligand. As expected from the K(d) value, the enzyme is reversibly inhibited upon exposure to pathologically, and possibly physiologically, relevant concentrations of NO. Inhibition of hHO-1 by NO may contribute to the pleiotropic responses to NO and CO.  相似文献   

8.
Heme oxygenases convert heme to free iron, CO, and biliverdin. Saccharomyces cerevisiae and Candida albicans express putative heme oxygenases that are required for the acquisition of iron from heme, a critical process for fungal survival and virulence. The putative heme oxygenases Hmx1 and CaHmx1 from S. cerevisiae and C. albicans, respectively, minus the sequences coding for C-terminal membrane-binding domains, have been expressed in Escherichia coli. The C-terminal His-tagged, truncated enzymes are obtained as soluble, active proteins. Purified ferric Hmx1 and CaHmx1 have Soret absorption maxima at 404 and 410 nm, respectively. The apparent heme binding Kd values for Hmx1 and CaHmx1 are 0.34 +/- 0.09 microM and 1.0 +/- 0.2 microM, respectively. The resonance Raman spectra of Hmx1 reveal a heme binding pocket similar to those of the mammalian and bacterial heme oxygenases. Several reductants, including ascorbate, yeast cytochrome P450 reductase (CPR), human CPR, spinach ferredoxin/ferredoxin reductase, and putidaredoxin/putidaredoxin reductase, are able to provide electrons for biliverdin production by Hmx1 and CaHmx1. Of these, ascorbate is the most effective reducing partner. Heme oxidation by Hmx1 and CaHmx1 regiospecifically produces biliverdin IXalpha. Spectroscopic analysis of aerobic reactions with H2O2 identifies verdoheme as a reaction intermediate. Hmx1 and CaHmx1 are the first fungal heme oxygenases to be heterologously overexpressed and characterized. Their heme degradation activity is consistent with a role in iron acquisition.  相似文献   

9.
Heme oxygenase oxidatively degrades heme to biliverdin resulting in the release of iron and CO through a process in which the heme participates both as a cofactor and substrate. One of the least understood steps in the heme degradation pathway is the conversion of verdoheme to biliverdin. In order to obtain a better understanding of this step we report the crystal structures of ferrous-verdoheme and, as a mimic for the oxy-verdoheme complex, ferrous-NO verdoheme in a complex with human HO-1 at 2.20 and 2.10 A, respectively. In both structures the verdoheme occupies the same binding location as heme in heme-HO-1, but rather than being ruffled verdoheme in both sets of structures is flat. Both structures are similar to their heme counterparts except for the distal helix and heme pocket solvent structure. In the ferrous-verdoheme structure the distal helix moves closer to the verdoheme, thus tightening the active site. NO binds to verdoheme in a similar bent conformation to that found in heme-HO-1. The bend angle in the verodoheme-NO structure places the terminal NO oxygen 1 A closer to the alpha-meso oxygen of verdoheme compared to the alpha-meso carbon on the heme-NO structure. A network of water molecules, which provide the required protons to activate the iron-oxy complex of heme-HO-1, is absent in both ferrous-verdoheme and the verdoheme-NO structure.  相似文献   

10.
Heme oxygenases catalyze the oxidation of heme to biliverdin, CO, and free iron. For pathogenic microorganisms, heme uptake and degradation are critical mechanisms for iron acquisition that enable multiplication and survival within hosts they invade. Here we report the first crystal structure of the pathogenic Escherichia coli O157:H7 heme oxygenase ChuS in complex with heme at 1.45 A resolution. When compared with other heme oxygenases, ChuS has a unique fold, including structural repeats and a beta-sheet core. Not surprisingly, the mode of heme coordination by ChuS is also distinct, whereby heme is largely stabilized by residues from the C-terminal domain, assisted by a distant arginine from the N-terminal domain. Upon heme binding, there is no large conformational change beyond the fine tuning of a key histidine (His-193) residue. Most intriguingly, in contrast to other heme oxygenases, the propionic side chains of heme are orientated toward the protein core, exposing the alpha-meso carbon position where O(2) is added during heme degradation. This unique orientation may facilitate presentation to an electron donor, explaining the significantly reduced concentration of ascorbic acid needed for the reaction. Based on the ChuS-heme structure, we converted the histidine residue responsible for axial coordination of the heme group to an asparagine residue (H193N), as well as converting a second histidine to an alanine residue (H73A) for comparison purposes. We employed spectral analysis and CO measurement by gas chromatography to analyze catalysis by ChuS, H193N, and H73A, demonstrating that His-193 is the key residue for the heme-degrading activity of ChuS.  相似文献   

11.
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   

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.
Heme oxygenase (HO) catalyzes physiological heme degradation consisting of three sequential oxidation steps that use dioxygen molecules and reducing equivalents. We determined the crystal structure of rat HO-1 in complex with heme and azide (HO-heme-N(3)(-)) at 1.9-A resolution. The azide, whose terminal nitrogen atom is coordinated to the ferric heme iron, is situated nearly parallel to the heme plane, and its other end is directed toward the alpha-meso position of the heme. Based on resonance Raman spectroscopic analysis of HO-heme bound to dioxygen, this parallel coordination mode suggests that the azide is an analog of dioxygen. The azide is surrounded by residues of the distal F-helix with only the direction to the alpha-meso carbon being open. This indicates that regiospecific oxygenation of the heme is primarily caused by the steric constraint between the dioxygen bound to heme and the F-helix. The azide interacts with Asp-140, Arg-136, and Thr-135 through a hydrogen bond network involving five water molecules on the distal side of the heme. This network, also present in HO-heme, may function in dioxygen activation in the first hydroxylation step. From the orientation of azide in HO-heme-N(3)(-), the dioxygen or hydroperoxide bound to HO-heme, the active oxygen species of the first reaction, is inferred to have a similar orientation suitable for a direct attack on the alpha-meso carbon.  相似文献   

14.
A major inducible form of heme oxygenase (EC 1.14.99.3) was purified from liver microsomes of chicks pretreated with cadmium chloride. The purification involved solubilization of microsomes with Emulgen 913 and sodium cholate, followed by DEAE-Sephacel, carboxymethyl-cellulose (CM-52) and hydroxyapatite chromatography, and FPLC through Superose 6 and 12 columns operating in series. The final product gave a single band on silver-stained SDS/polyacrylamide gels (Mr = 33,000). Optimal conditions for measurement of activity of solubilized heme oxygenase were studied. In a reconstituted system containing purified heme oxygenase, NADPH-cytochrome reductase, biliverdin reductase and NADPH, the Km for free heme was 3.8 +/- 0.5 microM; for heme in the presence of bovine serum albumin (5 mol heme/3 mol albumin) the Km was 5.0 +/- 0.8 microM; and the Km for NADPH was 6.1 +/- 0.4 microM (all values mean +/- SD, n = 3). Oxygen concentration as low as 15 microM, with saturating concentrations of heme and NADPH, did not affect the reaction rate, indicating that the supply of oxygen is not involved in the physiological regulation of activity of the enzyme. The pH optimum of the reaction was 7.4; at 37 degrees C, the apparent Vmax was 580 +/- 44 nmol biliverdin.(mg protein)-1.min-1 and the molecular activity was 19.2 min-1. Biliverdin IXa was the sole biliverdin isomer formed. In the presence of purified biliverdin reductase, biliverdin was converted quantitatively to bilirubin. Addition of catalase to the reconstituted system decreased the breakdown of heme to non-biliverdin products and led to nearly stoichiometric conversion of heme to biliverdin. Activity of the enzyme in the reconstituted system was inhibited by metalloporphyrins in the following order of decreasing potency: tin mesoporphyrin greater than tin protoporphyrin greater than zinc protoporphyrin greater than manganese protoporphyrin greater than cobalt protoporphyrin. Protoporphyrin (3.3 or 6.6 microM) (and several other porphyrins) and metallic ions (100 microM) alone had little if any inhibitory effect, except for Hg2+ which inhibited by 67% at 10 microM and totally at 15 microM. Following partial cleavage, fragments of the purified enzyme were sequenced. Comparison of sequences to those derived from cDNA sequences for the major inducible rat and human heme oxygenase showed 69% and 76% similarities, respectively. The histidine residue at position 132 of rat heme oxygenase-1 and the residues (Lys128-Arg136) flanking His132 were conserved in all three enzymes, as well as in the corresponding portion of a fourth less highly similar rat enzyme, heme oxygenase-2.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
16.
Carbon monoxide (CO) stimulates guanylate cyclase (GC) and increases guanosine 3',5'-cyclic monophosphate (cGMP) levels. We transfected rat-lung pulmonary endothelial cells with a retrovirus-mediated human heme oxygenase (hHO)-1 gene. Pulmonary cells that expressed hHO-1 exhibited a fourfold increase in HO activity associated with decreases in the steady-state levels of heme and cGMP without changes in soluble GC (sGC) and endothelial nitric oxide synthase (NOS) proteins or basal nitrite production. Heme elicited significant increases in CO production and intracellular cGMP levels in both pulmonary endothelial and pulmonary hHO-1-expressing cells. N(omega)-nitro-L-arginine methyl ester (L-NAME), an inhibitor of NOS, significantly decreased cGMP levels in heme-treated pulmonary endothelial cells but not heme-treated hHO-1-expressing cells. In the presence of exogenous heme, CO and cGMP levels in hHO-1-expressing cells exceeded the corresponding levels in pulmonary endothelial cells. Acute exposure of endothelial cells to SnCl2, which is an inducer of HO-1, increased cGMP levels, whereas chronic exposure decreased heme and cGMP levels. These results indicate that prolonged overexpression of HO-1 ultimately decreases sGC activity by limiting the availability of cellular heme. Heme activates sGC and enhances cGMP levels via a mechanism that is largely insensitive to NOS inhibition.  相似文献   

17.
We cloned a cDNA for a Drosophila melanogaster homologue of mammalian heme oxygenase (HO) and constructed a bacterial expression system of a truncated, soluble form of D. melanogaster HO (DmDeltaHO). The purified DmDeltaHO degraded hemin to biliverdin, CO and iron in the presence of reducing systems such as NADPH/cytochrome P450 reductase and sodium ascorbate, although the reaction rate was slower than that of mammalian HOs. Some properties of DmHO, however, are quite different from other known HOs. Thus DmDeltaHO bound hemin stoichiometrically to form a hemin-enzyme complex like other HOs, but this complex did not show an absorption spectrum of hexa-coordinated heme protein. The absorption spectrum of the ferric complex was not influenced by changing the pH of the solution. Interestingly, an EPR study revealed that the iron of heme was not involved in binding heme to the enzyme. Hydrogen peroxide failed to convert it into verdoheme. A spectrum of the ferrous-CO form of verdoheme was not detected during the reaction from hemin under oxygen and CO. Degradation of hemin catalyzed by DmDeltaHO yielded three isomers of biliverdin, of which biliverdin IXalpha and two other isomers (IXbeta and IXdelta) accounted for 75% and 25%, respectively. Taken together, we conclude that, although DmHO acts as a real HO in D. melanogaster, its active-site structure is quite different from those of other known HOs.  相似文献   

18.
M Tanaka  K Ishimori  I Morishima 《Biochemistry》1999,38(32):10463-10473
To enhance the oxidation activity for luminol in horseradish peroxidase (HRP), we have prepared three HRP mutants by mimicking a possible binding site for luminol in Arthromyces ramosus peroxidase (ARP) which shows 500-fold higher oxidation activity for luminol than native HRP. Spectroscopic studies by (1)H NMR revealed that the chemical shifts of 7-propionate and 8-methyl protons of the heme in cyanide-ligated ARP were deviated upon addition of luminol (4 mM), suggesting that the charged residues, Lys49 and Glu190, which are located near the 7-propionate and 8-methyl groups of the heme, are involved in the specific binding to luminol. The positively charged Lys and negatively charged Glu were introduced into the corresponding positions of Ser35 (S35K) and Gln176 (Q176E) in HRP, respectively, to build the putative binding site for luminol. A double mutant, S35K/Q176E, in which both Ser35 and Gln176 were replaced, was also prepared. Addition of luminol to the HRP mutants induced more pronounced effects on the resonances from the heme substituents and heme environmental residues in the (1)H NMR spectra than that to the wild-type enzyme, indicating that the mutations in this study induced interactions with luminol in the vicinity of the heme. The catalytic efficiencies (V(max)/K(m)) for luminol oxidation of the S35K and S35K/Q176E mutants were 1.5- and 2-fold improved, whereas that of the Q176E mutant was slightly depressed. The increase in luminol activity of the S35K and S35K/Q176E mutants was rather small but significant, suggesting that the electrostatic interactions between the positive charge of Lys35 and the negative charge of luminol can contribute to the effective binding for the luminol oxidation. On the other hand, the negatively charged residue would not be so crucial for the luminol oxidation. The absence of drastic improvement in the luminol activity suggests that introduction of the charged residues into the heme vicinity is not enough to enhance the oxidation activity for luminol as observed for ARP.  相似文献   

19.
Biliverdin IXbeta reductase (BVR-B) catalyzes the pyridine nucleotide-dependent production of bilirubin-IXbeta, the major heme catabolite during early fetal development. BVR-B displays a preference for biliverdin isomers without propionates straddling the C10 position, in contrast to biliverdin IXalpha reductase (BVR-A), the major form of BVR in adult human liver. In addition to its tetrapyrrole clearance role in the fetus, BVR-B has flavin and ferric reductase activities in the adult. We have solved the structure of human BVR-B in complex with NADP+ at 1.15 A resolution. Human BVR-B is a monomer displaying an alpha/beta dinucleotide binding fold. The structures of ternary complexes with mesobiliverdin IValpha, biliverdin IXalpha, FMN and lumichrome show that human BVR-B has a single substrate binding site, to which substrates and inhibitors bind primarily through hydrophobic interactions, explaining its broad specificity. The reducible atom of both biliverdin and flavin substrates lies above the reactive C4 of the cofactor, an appropriate position for direct hydride transfer. BVR-B discriminates against the biliverdin IXalpha isomer through steric hindrance at the bilatriene side chain binding pockets. The structure also explains the enzyme's preference for NADP(H) and its B-face stereospecificity.  相似文献   

20.
The crystal structure of rat heme oxygenase-1 in complex with biliverdin-iron chelate (biliverdin(Fe)-HO-1), the immediate precursor of the final product, biliverdin, has been determined at a 2.4-A resolution. The electron density in the heme pocket clearly showed that the tetrapyrrole ring of heme is cleaved at the alpha-meso edge. Like the heme bound to HO-1, biliverdin-iron chelate is located between the distal and proximal helices, but its accommodation state seems to be less stable in light of the disordering of the solvent-exposed propionate and vinyl groups. The middle of the distal helix is shifted away from the center of the active site in biliverdin(Fe)-HO-1, increasing the size of the heme pocket. The hydrogen-bonding interaction between Glu-29 and Gln-38, considered to restrain the orientation of the proximal helix in the heme-HO-1 complex, was lost in biliverdin(Fe)-HO-1, leading to relaxation of the helix. Biliverdin has a distorted helical conformation; the lactam oxygen atom of its pyrrole ring-A interacted with Asp-140 through a hydrogen-bonding solvent network. Because of the absence of a distal water ligand, the iron atom is five-coordinated with His-25 and four pyrrole nitrogen atoms. The coordination geometry deviates considerably from a square pyramid, suggesting that the iron may be readily dissociated. We speculate that the opened conformation of the heme pocket facilitates sequential product release, first iron then biliverdin, and that because of biliverdin's increased flexibility, iron release triggers its slow dissociation.  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