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 15 毫秒
1.
Site-specific mutants of human Cu,Zn superoxide dismutase (Cu,ZnSOD) have been prepared in which the active-site arginine at position 143 (i.e., SODR143) has been replaced by either lysine (SODK143) or isoleucine (SODI143). As reported previously (W.F. Beyer, Jr., et al. (1987) J. Biol. Chem. 262, 11182-11187), SODK143 and SODI143 have 43 and 11%, respectively, of the catalytic activity of SODR143. H2O2, at low concentrations, acts as an affinity reagent for the inactivation of SODR143. At pH 9.0 and 25 degrees C, the process is characterized by a half-saturation constant for H2O2, K50, of 5.1 mM and a maximum pseudo-first-order rate constant for inactivation, Kmax, of 0.53 min-1. At pH 11.5, the corresponding values are 0.63 mM and 1.23 min-1. The active species in the inactivation is likely HO2-, as previously found with yeast and bovine Cu,ZnSODs (see C.L. Borders, Jr., and I. Fridovich (1985) Arch. Biochem. Biophys. 241, 472-476). SODK143 is also inactivated by HO2- by an affinity mechanism, i.e., one where reversible binding of H2O2 (HO2-) is a prerequisite for inactivation. At pH values of 9.0 and 11.5, the kmax values are 0.92 and 1.08 min-1, respectively; however, the corresponding K50 values increase to 42.5 and 15.8 mM, respectively. SODI143 is also inactivated by H2O2, but no evidence for an affinity mechanism was found; instead, a second-order kinetic mechanism was observed. Inactivation of each of the three enzymes is accompanied by the loss of one histidine per subunit. At elevated concentrations of H2O2, a second nonaffinity mechanism of inactivation of both SODR143 and SODK143 was found, in which a second equivalent of H2O2 reacts with the Cu,ZnSOD.HO2- complex to give a competing second-order inactivation. It appears that the positive charge of arginine-143 plays a role in the binding of HO2- at the active site of human Cu,ZnSOD, and that replacement of the arginine by lysine gives an enzyme with a similar affinity mechanism of inactivation, but with a greatly reduced affinity for HO2-. However, replacement with isoleucine causes an entirely different mechanism of inactivation; this raises the possibility that the mechanism of enzyme catalysis of superoxide dismutation by SODI143 is also different.  相似文献   

2.
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   

3.
R van Wijk  L Loesberg  G T Snoek 《Biochimie》1983,65(11-12):643-652
Tyrosine aminotransferase activity increased during conversion of serum depleted quiescent Reuber H35 rat hepatoma cells into the proliferative state. Increased activity coincides with the actual increase of cells into S phase. The rate of tyrosine aminotransferase synthesis along the cell cycle was studied. The rate of enzyme synthesis fluctuated through the cell cycle but could not explain the increase of specific activity. Apparently enzyme activity is predominantly regulated by a post-translational event. Intracellular levels of cyclic AMP and cyclic GMP were measured at various times of G1 and S phases. In the early part of the cell cycle tyrosine aminotransferase decreased while intracellular levels of cyclic AMP increased. At later stages cyclic AMP rises concurrently with increased rates of enzyme synthesis. Induction of tyrosine aminotransferase by N6,O2'-dibutyryladenosine 3', 5'-monophosphate (Bt2cAMP) was studied. Inducibility by Bt2cAMP fluctuated through the cell cycle. Alternation of positive and negative control of tyrosine aminotransferase synthesis was observed. In early serum induced cells, Bt2cAMP increased enzyme activity without any increased rate of enzyme synthesis, on the contrary, a decreased rate of synthesis was observed. The data support the view that alternation of positive and negative control of tyrosine aminotransferase synthesis and temporary post-translational control of enzyme activity determine the enzyme level during the transition of quiescent hepatoma cells into proliferation.  相似文献   

4.
The claim that peroxidase (rather than tyrosinase) is the enzyme responsible for the conversion of tyrosine into dopa (3,4-dihydroxyphenylalanine) in melanogenesis was investigated. The spectral changes that occurred during the action of horseradish peroxidase in the presence of H2O2 on dopa, tyrosine and mixtures of dopa with tyrosine or other phenolic compounds were studied. The effect of ascorbic acid or dihydroxyfumaric acid on some of these changes was also investigated. No evidence was found that tyrosine was hydroxylated by peroxidase in the presence of H2O2 and dopa as cofactor, although tyrosine or other phenolic compounds increased the rate of oxidation of dopa to dopachrome (indoline-5,6-quinone-2-carboxylic acid). Peroxidase was, however, effective in oxidizing tyrosine to dopa in the presence of dihydroxyfumaric acid and oxygen.  相似文献   

5.
Two enzymes which transaminate tyrosine and phenylalanine in Bacillus subtilis were each purified over 200-fold and partially characterized. One of the enzymes, termed histidinol phosphate aminotransferase, is also active with imidazole acetyl phosphate as the amino group recipient. Previous studies have shown that mutants lacking this enzyme require histidine for growth. Mutants in the other enzyme termed aromatic aminotransferase are prototrophs. Neither enzyme is active on any other substrate involved in amino acid synthesis. The two enzymes can be distinguished by a number of criteria. Gel filtration analysis indicate the aromatic and histidinol phosphate aminotransferases have molecular weights of 63,500 and 33,000, respectively. Histidinol phosphate aminotransferase is heat-sensitive, whereas aromatic aminotransferase is relatively heat-stable, particularly in the presence of alpha-ketoglutarate. Both enzymes display typical Michaelis-Menten kinetics in their rates of reaction. The two enzymes have similar pH optima and employ a ping-pong mechanism of action. The Km values for various substrates suggest that histidinol phosphate aminotransferase is the predominant enzyme responsible for the transamaination reactions in the synthesis of tyrosine and phenylalanine. This enzyme has a 4-fold higher affinity for tyrosine and phenylalanine than does the aromatic aminotransferase. Competitive substrate inhibition was observed between tyrosine, phenylalanine, and histidinol phosphate for histidinol phosphate aminotransferase. The significance of the fact that an enzyme of histidine synthesis plays an important role in aromatic amino acid synthesis is discussed.  相似文献   

6.
Bacterial oxidation of polyethylene glycol.   总被引:13,自引:8,他引:5       下载免费PDF全文
The metabolism of polyethylene glycol (PEG) was investigated with a synergistic, mixed culture of Flavobacterium and Pseudomonas species, which are individually unable to utilize PEGs. The PEG dehydrogenase linked with 2,6-dichlorophenolindophenol was found in the particulate fraction of sonic extracts and catalyzed the formation of a 2,4-dinitrophenylhydrazine-positive compound, possibly an an aldehyde. The enzyme has a wide substrate specificity towards PEGs: from diethylene glycol to PEG 20,000 Km values for tetraethylene glycol (TEG), PEG 400, and PEG 6,000 were 11, 1.7, and 15 mM, respectively. The metabolic products formed from TEG by intact cells were isolated and identified by combined gas chromatography-mass spectrometry as triethylene glycol and TEG-monocarboxylic acid plus small amounts of TEG-dicarboxylic acid, diethylene glycol, and ethylene glycol. From these enzymatic and analytical data, the following metabolic pathway was proposed for PEG: HO(CH2CH2O)nCH2CH2OH leads to HO(CH2CH2O)nCH2CHO leads to HO(CH2CH2O)nCH2COOH leads to HO(CH2CH2O)n-1CH2CH2OH.  相似文献   

7.
A peptide screened from a combinatorial peptide library with the sequence EYKSWEYC performed best as a ligand for affinity chromatography of human blood coagulation factor VIII (FVIII). With this peptide immobilized on monolithic CIM columns via epoxy groups we were able to capture FVIII from diluted plasma. Rational substitution of amino acids by spot synthesis revealed that lysine and cysteine can be exchanged for almost all other proteinogenic amino acids without loss of affinity to FVIII. This offers the possibility of site-specific attachment via either one of these residues or the N- or C-terminus. The aliphatic positions O5 (tryptophan) and O7 (tyrosine), together with the charged position O6 (glutamic acid), seem to form the core of the binding unit. In the positions with aliphatic amino acids, substitution by tyrosine or phenylalanine, and in the positions with charged amino acids, substitution by aspartic acid or lysine, preserved the affinity to FVIII. The functionality of the selected peptides was confirmed by affinity chromatography. Selective binding and elution could be achieved.  相似文献   

8.
The Mg(2+) cofactor of the F(1)F(0) ATP synthase is required for the asymmetry of the catalytic sites that leads to the differences in affinity for nucleotides. Vanadyl (V(IV)=O)(2+) is a functional surrogate for Mg(2+) in the F(1)-ATPase. The (51)V-hyperfine parameters derived from EPR spectra of VO(2+) bound to specific sites on the enzyme provide a direct probe of the metal ligands at each site. Site-directed mutations of residues that serve as metal ligands were found to cause measurable changes in the (51)V-hyperfine parameters of the bound VO(2+), thereby providing a means by which metal ligands were identified in the functional enzyme in several conformations. At the low-affinity catalytic site comparable to beta(E) in mitochondrial F(1), activation of the chloroplast F(1)-ATPase activity induces a conformational change that inserts the P-loop threonine and catch-loop tyrosine hydroxyl groups into the metal coordination sphere thereby displacing an amino group and the Walker homology B aspartate. Kinetic evidence suggests that coordination of this tyrosine by the metal when the empty site binds substrate may provide an escapement mechanism that allows the gamma subunit to rotate and the conformation of the catalytic sites to change, thereby allowing rotation only when the catalytic sites are filled. In the high-affinity conformation analogous to the beta(DP) site of mitochondrial F(1), the catch-loop tyrosine has been displaced by carboxyl groups from the Walker homology B aspartate and from betaE197 in Chlamydomonas CF(1). Coordination of the metal by these carboxyl groups contributes significantly to the ability of the enzyme to bind the nucleotide with high affinity.  相似文献   

9.
N R Matheson  J Travis 《Biochemistry》1985,24(8):1941-1945
Human alpha 1-proteinase inhibitor is easily susceptible to inactivation because of the presence of a methionyl residue at its reactive site. Thus, oxidizing species derived from the myeloperoxidase system (enzyme, H2O2, and C1-), as well as hypochlorous acid, can inactivate this inhibitor, although H2O2 alone has no effect. Butylated hydroxytoluene, a radical scavenger, partially protects alpha 1-proteinase inhibitor from the myeloperoxidase system and completely protects it from hypochlorous acid. Each oxidant also reacts differently with the inhibitor, in that the myeloperoxidase system and hypochlorous acid can each oxidize as many as six methionyl residues, but hypochlorous acid can also oxidize a single tyrosine residue. Myeloperoxidase can be inactivated by hypochlorous acid, by autoxidation in the presence of H2O2 and C1-, as well as by H2O2 alone. Butylated hydroxytoluene completely protects this enzyme from hypochlorous acid inactivation, does not affect the action of H2O2, and enhances autoinactivation. As many as six methionyl residues and two tyrosine residues could be oxidized during autoxidation and six methionine residues by H2O2 alone. Eight methionine residues and one tyrosine residue could be oxidized by hypochlorous acid. The tyrosine residue in myeloperoxidase was oxidized only at a relatively high concentration (600 microM) of hypochlorous acid at which point the enzyme simultaneously and completely lost its enzymatic activity. Loss of activity of myeloperoxidase could also be correlated with the loss of the heme groups present in the enzyme when a relatively high concentration of hypochlorous acid (600 microM) was used and also during autoxidation. It appears that once there is sufficient oxidant to modify one of the tyrosine residues, the heme group itself becomes susceptible.  相似文献   

10.
The reaction of hydrogen peroxide with the copper-zinc bovine-liver superoxide dismutase at low molar ratios (0.2-20.0) of H2O2/active site between pH 7.3-10.0 leads to the loss of native enzyme as a distinct form monitored by electrophoresis. The pH dependence of the loss of native enzyme between 7.3 and 9.0 indicates the involvement of a conjugate base on the enzyme of pKa of 8.7 +/- 0.1. The rate of loss of the native enzyme is first order with respect to the concentration of both enzyme and hydrogen peroxide between pH 7.3 and 9.0 with no evidence for binding of peroxide. A second-order rate constant of 3.0 +/- 1.0 M-1 s-1 is obtained from these data. At pH 10.0 the reaction is first order with respect to enzyme concentration but saturable in H2O2. All data are consistent with the interpretation that H2O2 reacts with the enzyme at the lower pH where the reaction is dependent upon the conjugate base of a functional group on the enzyme. At the higher pH, the data are consistent with the reaction of HO2- and H2O2 with the dismutase. The dissociation constant for HO2- calculated from the kinetic data at pH 10.0 is between 25-50 microM and the rate constant for the breakdown of the HO2- dismutase complex is 1.10 + 0.05 x 10(-2) s-1. The change in the electrophoretic pattern at all pH values is accompanied by the loss of the ability of the enzyme to bind copper. Weakly bound or free copper can be detected using bathocuproine disulfonate. Furthermore copper-defficient forms of the enzyme can be detected by staining gels of the peroxide-treated dismutase with diethyldithiocarbamate.  相似文献   

11.
C Auclair  M Torres  E Cramer  J Hakim 《Enzyme》1978,23(4):225-237
The biochemical triad, NADH oxidation, oxygen (O2) uptake and hydrogen peroxide (H2O2) formation, by subcellular fractions of human blood polymorphonuclears (PMNs) was investigated. It was found that this biochemical triad (1) was under the control of the granule-rich fraction (GRF) only; (2) was not inhibited by cyanide; (3) occurred stoichiometrically for its three components, and (4) accounted quantitatively for the respiratory burst of the stimulated PMN. It was also shown that the above biochemical triad (1) involved an enzymatic step; (2) was enhanced by acidic pH (0.5) and Mg++; (3) was inhibited by Cu++ or low concentration of Mn++; (4) was dependent on H2O2, perhydroxyl radical (HO2) and hydroxyl radical (HO) since either catalase or superoxide dismutase or scavengers of HO2 or HO were inhibitor, and (5) involved multistep reactions. Evidence is provided that the sequence of the reactions is first a generation of H2O2, (spontaneously from NADH in our incubation medium), secondly the production of HO from H2O2, thirdly the oxidation of NADH with further production of HO2,O2 uptake and H2O2 formation, probably through a chain reaction. The identification of the enzyme(s) involved in these multistep reactions needs further studies.  相似文献   

12.
4-Hydroxyphenylacetic acid 3-hydroxylase is a key enzyme in the pathway for the microbial degradation of phenylalanine, tyrosine and many aromatic amines. This enzyme was purified to homogeneity from Pseudomonas putida by affinity chromatography. The protein had a molecular weight of 91,000 and was a dimer of identical subunits. It was a typical external flavoprotein monooxygenase and showed an absolute requirement of NADH for activity. The enzyme had a pH optimum of 7.5 and the Km values for 4-hydroxyphenylacetic acid and NADH were 2 x 10(-4) M and 5.9 x 10(-5) M respectively. It was strongly inhibited by heavy metal ions and thiol reagents, suggesting the possible involvement of -SH group(s) in enzyme reaction.  相似文献   

13.
The hydroxyl radical (HO*)- and superoxide anion radical (O* (2))-scavenging activity, as well as the singlet oxygen ((1)O(2))-quenching property of N-substituted indole-2-carboxylic acid esters (INDs) were investigated by deoxyribose degradation assay, a chemiluminescence method and the electron spin resonance (ESR) spin-trapping technique. This novel group of compounds was developed as a search for cyclooxygenase-2 (COX-2)-selective enzyme inhibitors. The results obtained demonstrated that of the 16 compounds examined, five inhibited light emission from the superoxide anion radical (O* (2))-DMSO system by at least 60% at a concentration of 1 mmol/L, nine prevented the degradation of deoxyribose induced by the Fenton reaction system (range 3-78%) or scavenged hydroxyl radicals (HO*) directly (range 8-93%) and 14 showed the (1)O(2)-quenching effect (range 10-74%). These results indicate that majority of the indole esters tested possess the ability to scavenge O(-) (2) and HO radicals and to quench (1)O(2) directly, and consequently may be considered effective antioxidative agents.  相似文献   

14.
Here, the PduX enzyme of Salmonella enterica is shown to be an L-threonine kinase used for the de novo synthesis of coenzyme B(12) and the assimilation of cobyric acid (Cby). PduX with a C-terminal His tag (PduX-His(6)) was produced at high levels in Escherichia coli, purified by nickel affinity chromatography, and partially characterized. (31)P NMR spectroscopy established that purified PduX-His(6) catalyzed the conversion of l-threonine and ATP to L-threonine-O-3-phosphate and ADP. Enzyme assays showed that ATP was the preferred substrate compared with GTP, CTP, or UTP. PduX displayed Michaelis-Menten kinetics with respect to both ATP and l-threonine and nonlinear regression was used to determine the following kinetic constants: V(max) = 62.1 +/- 3.6 nmol min(-1) mg of protein(-1); K(m)(, ATP) = 54.7 +/- 5.7 microm and K(m)(,Thr) = 146.1 +/- 8.4 microm. Growth studies showed that pduX mutants were impaired for the synthesis of coenzyme B(12) de novo and from Cby, but not from cobinamide, which was the expected phenotype for an L-threonine kinase mutant. The defect in Cby assimilation was corrected by ectopic expression of pduX or by supplementation of growth medium with L-threonine-O-3-phosphate, providing further support that PduX is an L-threonine kinase. In addition, a bioassay showed that a pduX mutant was impaired for the de novo synthesis of coenzyme B(12) as expected. Collectively, the genetic and biochemical studies presented here show that PduX is an L-threonine kinase used for AdoCbl synthesis. To our knowledge, PduX is the first enzyme shown to phosphorylate free L-threonine and the first L-threonine kinase shown to function in coenzyme B(12) synthesis.  相似文献   

15.
The following compounds were tested as early intermediates in the conversion of tyrosine to p-hydroxymandelonitrile by a microsomal preparation from dark grown sorghum seedlings: p-hydroxyphenylacetamide, 1-nitro-2-p-hydroxyphenylethane, p-hydroxyphenyl-pyruvic acid oxime, tyramine, N-hydroxytyramine, and N-hydroxytyrosine. Of these, only N-hydroxytyrosine was metabolized to p-hydroxymandelonitrile. N-Hydroxytyrosine was produced from L-[U-14C]tyrosine in tracer experiments when unlabeled N-hydroxytyrosine was added as a trap. These data indicate N-hydroxytyrosine as the first intermediate in the biosynthesis of dhurrin, the cyanogenic glucoside of sorghum, and represent the first demonstration of the formation of an alpha-N-hydroxy-amino acid in a biological system. The enzyme system involved in this reaction was partially characterized with respect to substrate specificity and the effect of various inhibitors. The enzyme was shown to have properties different than those reported for the mammalian enzyme system(s) involved in the N-hydroxylation of amine drugs. The possible involvement of N-hydroxyamino acids in the biosynthesis of other secondary plant products is discussed.  相似文献   

16.
The involvement of reactive oxygen species (ROS) in cardiac ischemia-reperfusion injuries is well-established, but the deleterious effects of hydrogen peroxide (H(2)O(2)), hydroxyl radical (HO*) or superoxide anion (O(2)*(-) ) on mitochondrial function are poorly understood. Here, we report that incubation of rat heart mitochondria with each of these three species resulted in a decline of the ADP-stimulated respiratory rate but not substrate-dependent respiration. These three species reduced oxygen consumption induced by an uncoupler without alteration of the respiratory chain complexes, but did not modify mitochondrial membrane permeability. HO* slightly decreased F1F0-ATPase activity and HO* and O(2)*(-) partially inhibited the activity of adenine nucleotide translocase; H(2)O(2) failed to alter these targets. They inhibited NADH production by acting specifically on aconitase for O(2)*(-) and alpha-ketoglutarate dehydrogenase for H(2)O(2) and HO*. Our results show that O(2)*(-), H(2)O(2) and HO* act on different mitochondrial targets to alter ATP synthesis, mostly through inhibition of NADH production.  相似文献   

17.
The in silico analysis of the amino acid sequences deduced from the complete genome sequence of Staphylococcus aureus suggested the presence of two protein tyrosine kinase activities, each split into two distinct polypeptides, respectively Cap5A1/Cap5B1 and Cap5A2/Cap5B2, like in some other Gram-positive bacteria. To check this prediction, the corresponding genes were cloned and overexpressed, and the four corresponding proteins were purified by affinity chromatography and assayed for phosphorylating activity in vitro. Individually, none of them was found to autophosphorylate. However, when Cap5B2 was incubated in the presence of Cap5A2 or, with a larger efficiency, in the presence of Cap5A1, this protein exhibited intensive autokinase activity, occurring selectively at tyrosine residues. On the other hand, whatever the protein combination assayed, Cap5B1 did not present any phosphorylating activity. In search of a possible role for the phosphorylation reaction mediated by Cap5B2, an endogenous substrate of this kinase was characterized. This substrate, termed Cap5O, is the enzyme UDP-acetyl-mannosamine dehydrogenase involved in the cascade of reactions leading to the synthesis of the bacterial capsule. It represents the first endogenous substrate for a tyrosine kinase activity so far identified in S. aureus. The analysis of its dehydrogenase activity showed that it was positively controlled by its phosphorylation at tyrosine.  相似文献   

18.
The activity of the first enzyme of aromatic path 3-deoxy-D-arabino-heptuloso-7-phosphate-synthase (DAHP-synthase) is regulated by retro-inhibition and is a subject of repression. Analysis of partially purified preparations of the enzyme has revealed three isoenzymes: DAHP-synthase-Tyr, DAHP-synthase-Trp and DAHP-synthase-Phe, each of them being regulated by a corresponding amino acid. DAHP-synthase-Phe is a dominant isoenzyme presenting 70% of the enzyme activity, 30% inhibition of which is possible by 7.0 mkM of phenylalanine. DAHP-synthase-Tyr and DAHP-synthase-Trp are minor isoenzymes (sharing 15% of enzyme activity each) and are controlled by tyrosine and tryptophane correspondingly. 50% of inhibition of activity is possible by adding 0.7 and 0.8 mkM of corresponding amino acid. Regulation of the enzyme synthesis was studied in the Trp-, Phe- and Tyr- mutants. The enzyme activity was registered under the conditions of limiting and surplus of each aromatic amino acid. The synthesis of DAHP-synthase in M. flagellatum KT is repressed by tryptophane and tyrosine decreasing the synthesis 18.8 and 15.6 fold.  相似文献   

19.
Chemical modification of PABA synthase.   总被引:1,自引:0,他引:1  
p-Aminobenzoic acid (PABA) synthase catalyses the first step in folic acid biosynthesis, the conversion of chorismate to p-aminobenzoate. In general, difficulties in purification have permitted only limited investigation of this enzyme. However, in an attempt to identify possible active site residues, the E. coli enzyme has been incubated with a range of protein modifying agents. Results indicate that cysteine, histidine, arginine and tyrosine residues are important for enzyme activity. Attempts were made to determine the subunits upon which these residues were located.  相似文献   

20.
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   

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