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 224 毫秒
1.
The effect of pH on the kinetic parameters (Kms, Vs) of the reaction of adrenaline and Fe(II) (More's salt) oxidation by ceruloplasmin isolated from human donor blood was investigated. It was assumed that the imidazole group of histidine is functionally important for the above reactions. For Fe(II) the effect of the ionizeable group was observed during substrate binding to the ceruloplasmin molecule, whereas in the course of the adrenaline oxidation reaction it manifests itself during catalytic interaction of the substrate with the enzyme. The organic substrate can bind both to the protonated and to the non-protonated form of the enzyme. Fe(II) interacts only with the protonated form of the protein. In both cases, the rate-limiting step of the oxidase reaction is preceded by a single step, i.e., proton binding. The schemes describing the order of proton attachment in the course of the above reactions are proposed.  相似文献   

2.
The binding of Fe2+ to both apo and holo mammalian ferritin has been investigated under anaerobic conditions as a function of pH. In the pH range 6.0-7.5, 8.0 +/- 0.5 Fe2+ ions bind to each apoferritin molecule, but above pH 7.5, a pH-dependent Fe2+ binding profile is observed with up to 80 Fe2+ ions binding at pH 10.0. This Fe2+ binding is reversible and is accompanied by up to two H+ being released per Fe2+ bound at pH 10.0. The Fe2+ binding to apoferritin probably occurs in the 3-fold channels. A much larger and more complex pH-dependent Fe2+ binding stoichiometry was observed for holoferritin with up to 300 Fe2+ ions binding at pH 10.0. This pH-dependent Fe2+ binding was interpreted as Fe2+ interaction at the FeOOH mineral surface with displacement of H+ from -OH or phosphate surface groups by the incoming Fe2+ ions. Mossbauer spectroscopic measurements using 57Fe-labeled Fe2+ under anaerobic conditions showed that 57Fe2+ binding to holoferritin was accompanied by electron transfer to the core, yielding 57Fe3+, presumably bound to the mineral surface. Removal of added iron by Fe2+-specific chelating agents yielded 57Fe2+, demonstrating the reversibility of this electron-transfer process. The Fe2+ bound to apo- and holoferritin is readily converted to Fe3+ by exposure to O2 and strongly retained by the respective ferritin species.  相似文献   

3.
Iron deposition within the iron storage protein ferritin involves a complex series of events consisting of Fe(2+) binding, transport, and oxidation at ferroxidase sites and mineralization of a hydrous ferric oxide core, the storage form of iron. In the present study, we have examined the thermodynamic properties of Fe(2+) binding to recombinant human H-chain apoferritin (HuHF) by isothermal titration calorimetry (ITC) in order to determine the location of the primary ferrous ion binding sites on the protein and the principal pathways by which the Fe(2+) travels to the dinuclear ferroxidase center prior to its oxidation to Fe(3+). Calorimetric titrations show that the ferroxidase center is the principal locus for Fe(2+) binding with weaker binding sites elsewhere on the protein and that one site of the ferroxidase center, likely the His65 containing A-site, preferentially binds Fe(2+). That only one site of the ferroxidase center is occupied by Fe(2+) implies that Fe(2+) oxidation to form diFe(III) species might occur in a stepwise fashion. In dilute anaerobic protein solution (3-5 microM), only 12 Fe(2+)/protein bind at pH 6.51 increasing to 24 Fe(2+)/protein at pH 7.04 and 7.5. Mutation of ferroxidase center residues (E62K+H65G) eliminates the binding of Fe(2+) to the center, a result confirming the importance of one or both Glu62 and His65 residues in Fe(2+) binding. The total Fe(2+) binding capacity of the protein is reduced in the 3-fold hydrophilic channel variant S14 (D131I+E134F), indicating that the primary avenue by which Fe(2+) gains access to the interior of ferritin is through these eight channels. The binding stoichiometry of the channel variant is one-third that of the recombinant wild-type H-chain ferritin whereas the enthalpy and association constant for Fe(2+) binding are similar for the two with an average values (DeltaH degrees = 7.82 kJ/mol, binding constant K = 1.48 x 10(5) M(-)(1) at pH 7.04). Since channel mutations do not completely prevent Fe(2+) binding to the ferroxidase center, iron gains access to the center in approximately one-third of the channel variant molecules by other pathways.  相似文献   

4.
Listeria innocua Dps (DNA binding protein from starved cells) affords protection to DNA against oxidative damage and can accumulate about 500 iron atoms within its central cavity through a process facilitated by a ferroxidase center. The chemistry of iron binding and oxidation in Listeria Dps (LiDps, formerly described as a ferritin) using H(2)O(2) as oxidant was studied to further define the mechanism of iron deposition inside the protein and the role of LiDps in protecting DNA from oxidative damage. The relatively strong binding of 12 Fe(2+) to the apoprotein (K(D) approximately 0.023 microM) was demonstrated by isothermal titration calorimetry, fluorescence quenching, and pH stat experiments. Hydrogen peroxide was found to be a more efficient oxidant for the protein-bound Fe(2+) than O(2). Iron(II) oxidation by H(2)O(2) occurs with a stoichiometry of 2 Fe(2+)/H(2)O(2) in both the protein-based ferroxidation and subsequent mineralization reactions, indicating complete reduction of H(2)O(2) to H(2)O. Electron paramagnetic resonance (EPR) spin-trapping experiments demonstrated that LiDps attenuates the production of hydroxyl radical by Fenton chemistry. DNA cleavage assays showed that the protein, while not binding to DNA itself, protects it against the deleterious combination of Fe(2+) and H(2)O(2). The overall process of iron deposition and detoxification by LiDps is described by the following equations. For ferroxidation, Fe(2+) + Dps(Z)--> [(Fe(2+))-Dps](Z+1) + H(+) (Fe(2+) binding) and [(Fe(2+))-Dps](Z+1) + Fe(2+) + H(2)O(2) --> [(Fe(3+))(2)(O)(2)-Dps](Z+1) + 2H(+) (Fe(2+) oxidation/hydrolysis). For mineralization, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(O)OH((core)) + 4H(+) (Fe(2+) oxidation/hydrolysis). These reactions occur in place of undesirable odd-electron redox processes that produce hydroxyl radical.  相似文献   

5.
In order to survive in an oxygen environment, aerobic organisms have developed numerous mechanisms to protect against oxygen radicals and singlet oxygen. One such mechanism, which appears to have attained particular significance during primate evolution, is the direct scavenging of oxygen radicals, singlet oxygen, oxo-haem oxidants and hydroperoxyl radicals by uric acid. In the present paper we demonstrate that another important 'antioxidant' property of uric acid is the ability to form stable co-ordination complexes with iron ions. Formation of urate-Fe3+ complexes dramatically inhibits Fe3+-catalysed ascorbate oxidation, as well as lipid peroxidation in liposomes and rat liver microsomal fraction. In contrast with antioxidant scavenger reactions, the inhibition of ascorbate oxidation and lipid peroxidation provided by urate's ability to bind iron ions does not involve urate oxidation. Association constants (Ka) for urate-iron ion complexes were determined by fluorescence-quenching techniques. The Ka for a 1:1 urate-Fe3+ complex was found to be 2.4 X 10(5), whereas the Ka for a 1:1 urate-Fe2+ complex was determined to be 1.9 X 10(4). Our experiments also revealed that urate can form a 2:1 complex with Fe3+ with an association constant for the second urate molecule (K'a) of approx. 4.5 X 10(5). From these data we estimate an overall stability constant (Ks approximately equal to Ka X K'a) for urate-Fe3+ complexes of approx. 1.1 X 10(11). Polarographic measurements revealed that (upon binding) urate decreases the reduction potential for the Fe2+/Fe3+ half-reaction from -0.77 V to -0.67 V. Thus urate slightly diminishes the oxidizing potential of Fe3+. The present results provide a mechanistic explanation for our previous report that urate protects ascorbate from oxidation in human blood. The almost saturating concentration of urate normally found in human plasma (up to 0.6 mM) represents 5-10 times the plasma ascorbate concentration, and is orders of magnitude higher than the 'free' iron ion concentration. These considerations point to the physiological significance of our findings.  相似文献   

6.
Binding of thiocyanate and cyanide ions to Mn(III) protoporphyrin-apohorseradish peroxidase complex [Mn(III)HRP] was investigated by relaxation rate measurements (at 50.68 MHz) of 15N resonance of SC15N- and C15N-. At pH = 4.0 the apparent dissociation constant (KD) for thiocyanate and cyanide binding to Mn(III)HRP was deduced to be 156 and 42 mM, respectively. The pH dependence of the 15N line width as well as apparent dissociation constant for thiocyanate and cyanide binding were quantitatively analyzed on the basis of a reaction scheme in which thiocyanate and cyanide in deprotonated form bind to the enzyme in a protonated form. The binding of thiocyanate and cyanide to Mn(III)HRP was found to be facilitated by protonation of an ionizable group on the enzyme [Mn(III)HRP] with a pKa = 4.0. From competitive binding studies it was shown that iodide, thiocyanate and cyanide bind to Mn(III)HRP at the same site; however, the binding site for resorcinol is different. The apparent dissociation constant for iodide binding deduced from competitive binding studies was found to be 117 mM, which agrees very well with the iodide binding to ferric HRP. The binding of thiocyanate and cyanide was shown to be away from the metal center and the distance of the 15N of thiocyanate and cyanide from the paramagnetic manganese ion in Mn(III)HRP was found to be 6.9 and 6.6 A, respectively. Except for cyanide binding, these observations parallel with the iodide and thiocyanate ion binding to native Fe(III)HRP. Water proton relaxivity measurements showed the presence of a coordinated water molecule to Mn(III)HRP with the distance of Mn-H2O being calculated to be 2.6 A. The slow reactivity of H2O2 towards Mn(III)HRP could be attributed to the presence of water at the sixth coordination position of the manganese ion.  相似文献   

7.
Kinetic study of the oxidase reaction of ceruloplasmin   总被引:1,自引:0,他引:1  
The effects of ionic strength, buffer composition and pH on the oxidase activity of ceruloplasmin isolated from human donor blood were studied. The steady-state kinetics of ceruloplasmin-catalyzed oxidation of organic substrates (pyrocatechine, adrenaline, rho-phenyldiamine) and Fe(II) was analyzed. The relationship between the initial reaction rate and Fe(II) concentration is described by the Michaelis--Menten kinetics, that for organic substrates--by substrate activation or by a scheme which implicates the existence of two catalytic centers in the enzyme molecule. The inhibition of adrenaline oxidation by its reaction product, adrenochrome, is competitive. The reactions studied were shown to occur via the formation of a ternary complex.  相似文献   

8.
Of the two mitochondrial enzymes of the urea cycle, carbamoyl phosphate synthetase (CPS) was and ornithine transcarbamylase (OTC) was not inactivated by the Fe3+-oxygen-ascorbate model system for mixed-function oxidation [R. L. Levine, (1983) J. Biol. Chem. 258, 11828-11833]. The susceptibility of OTC was not increased by its substrates, products, or inhibitors, whereas that of CPS was markedly increased by acetylglutamate (its allosteric activator) when ATP was absent. Thus, acetylglutamate binds in the absence of ATP and exposes to oxidation essential groups of the enzyme. We estimate for this binding a KD value of 1.6 mM, which greatly exceeds the KD values (less than 10 microM) determined in the presence of ATP and bicarbonate. ATP, and even more, mixtures of ATP and bicarbonate protected CPS from inactivation. Acetylglutamate exposes the site for the ATP molecule that yields Pi, and it appears that ATP protects by binding at this site. Experiments of limited proteolysis with elastase suggest that oxidation prevents this binding of ATP and show that it accelerates cleavage of CPS by the protease, thus supporting the idea that oxidation may precede proteolysis. Trypsin, chymotrypsin, and papain also hydrolyze the oxidized enzyme considerably faster than the native enzyme. Our results also support the idea that oxidative inactivation is site specific and requires sites on the enzyme for Me2+ and, possibly, for a nucleotide.  相似文献   

9.
Chicken liver lactate dehydrogenase (L-lactate : NAD+ oxidoreductase, EC 1.1.1.27) irreversibly catalyses the oxidation of glyoxylate (hydrated form) (I) to oxalate (pH = 9.6) and the reduction of (non-hydrated form) (II) to glycolate (pH = 7.4). (I) attaches to the enzyme in the pyruvate binding site and (II) attaches to the enzyme at the L-lactate binding site. The oxidation of (I) (pH = 9.6) is adapted to the following mechanism: (see book). The abortive complexes, E-NADH-I and E-NAD+-II, are responsible for the inhibition by excess substrate in the reduction and oxidation systems, respectively. When lactate dehydrogenase and NAD+ are preincubated, E-NAD+- NAD+ appears and causes inhibition by excess NAD+ in the glyoxylate-lactate dehydrogenase-NAD+ and L-lactate-lactate dehydrogenase-NAD+ systems; the second NAD+ molecule attaches to the enzyme at the L-lactate binding site.  相似文献   

10.
Ferritins are nearly ubiquitous iron storage proteins playing a fundamental role in iron metabolism. They are composed of 24 subunits forming a spherical protein shell encompassing a central iron storage cavity. The iron storage mechanism involves the initial binding and subsequent O2-dependent oxidation of two Fe2+ ions located at sites A and B within the highly conserved dinuclear "ferroxidase center" in individual subunits. Unlike animal ferritins and the heme-containing bacterioferritins, the Escherichia coli ferritin possesses an additional iron-binding site (site C) located on the inner surface of the protein shell close to the ferroxidase center. We report the structures of five E. coli ferritin variants and their Fe3+ and Zn2+ (a redox-stable alternative for Fe2+) derivatives. Single carboxyl ligand replacements in sites A, B, and C gave unique effects on metal binding, which explain the observed changes in Fe2+ oxidation rates. Binding of Fe2+ at both A and B sites is clearly essential for rapid Fe2+ oxidation, and the linking of FeB2+ to FeC2+ enables the oxidation of three Fe2+ ions. The transient binding of Fe2+ at one of three newly observed Zn2+ sites may allow the oxidation of four Fe2+ by one dioxygen molecule.  相似文献   

11.
The pH dependence of steady-state parameters for aldehyde reduction and alcohol oxidation were determined in the human liver aldehyde reductase reaction. The maximum velocity of aldehyde reduction with NADPH or 3-acetyl pyridine adenine dinucleotide phosphate (3-APADPH) was pH independent at low pH but decreased at high pH with a pK of 8.9-9.6. The V/K for both nucleotides decreased below a pK of 5.7-6.2, as did the pKi of competitive inhibitors NADP and ATP-ribose, suggesting that the 2'-phosphate of the nucleotide has to be deprotonated for binding to the enzyme. The pK of the 2'-phosphate of NADPH appears to be perturbed in the ternary complexes to 5.2-5.4. The V/K for NADPH, the V/K for 3-APADPH, and the pKi of ATP-ribose also decreased above a pK of 9-10, suggesting interaction of the 2'-phosphate of the nucleotide with a protonated base, perhaps lysine. Since protonation of a residue with a pK of 8 (evident in V/K for DL-glyceraldehyde and V/K for L-gulonate versus pH profiles) appears to be essential for aldehyde reduction, and deprotonation for alcohol oxidation, this residue appears to act as a general acid-base catalyst. An additional anion binding site with a pK of 9.94 facilitates the binding of carboxylic substrates such as D-glucuronate. With NADPH as the coenzyme the primary deuterium isotope effects on V and V/K for NADPH were close to unity and pH independent, suggesting that the hydride transfer step is not rate determining over the experimental pH range. With 3-APADPH as the coenzyme, the maximum velocity, relative to NADPH was three- to four-fold lower. Isotope effects on V, V/K for 3-APADPH, and V/K for D-glucuronate were pH independent and equal to 2.2-2.8, indicating that the chemical step of the reaction is relatively insensitive to pH. These data suggest that substrates bind to both the protonated and the deprotonated forms of the enzyme, though only the protonated enzyme catalyzes aldehyde reduction and the deprotonated enzyme catalyzes alcohol oxidation. On the basis of these results a scheme for the chemical mechanism of aldehyde reductase is postulated.  相似文献   

12.
Tryptamine, serotonin and tryptophan are readily oxidized during the Cu2+-catalyzed peroxidation of arachidonic acid (AA) at neutral pH and under certain experimental conditions which determine their relative susceptibility to oxidation. Thus, in AA micelles, fluorescence spectroscopy demonstrates that positively-charged indoles interact with negatively-charged micelles while Trp remains in the aqueous phase. As a result, serotonin and tryptamine are preferentially oxidized. In egg phosphatidylcholine liposomes loaded with AA, the three substrates interact with vesicles and undergo lipid-induced oxidation. EDTA inhibits the formation of thiobarbituric-reactive substances (TBARS) and prevents the indoles from oxidation. Owing to the intricate contact between the lipidic core and the apolipoproteins, the Trp residues of human serum LDL and HDL3 are very rapidly oxidized, i.e., at least one order of magnitude faster than Tyr HDL and Lys LDL, which are believed to be involved in the binding of these lipoproteins to their cell receptors. Cupric ions are rather specific for the lipid-induced autoxidation of Trp residues of lipoproteins whereas in micelles and liposomes, Mn2+ and Fe2+ can lead to TBARS production and to oxidation of indoles. This specificity is surprising considering the known ability of Fe2+ to catalyze LDL modification (measured by TBARS production) during their incubation with various cells. Biological consequences of the easy lipid-induced oxidation of biologically important indoles are discussed.  相似文献   

13.
Tris and two of its hydroxylated amine analogs were examined in a metal-free, universal n-butylamine buffer, for their interaction with intestinal brush border sucrase. Our recent three-proton-families model (Vasseur, van Melle, Frangne and Alvarado (1988) Biochem. J., 251, 667-675) has provided the sucrase pK values necessary to interpret the present work. At pH 5.2, 2-amino-2-methyl-l-propanol (PM) causes activation whereas Tris has a concentration-dependent biphasic effect, first causing activation, then fully competitive inhibition. The amine species causing activation is the protonated, cationic form. The difference between the two amines is related to the fact that Tris has a much lower pKa value than PM (respectively, 8.2 and 9.8). Even at pH 5.2, Tris (but not PM) exists as a significant proportion of the free base which, by inhibiting the enzyme fully competitively, overshadows the activating effect of the cationic, protonated amine. Above pH 6.8, both Tris and PM act as fully competitive inhibitors. These inhibitions increase monotonically between pH 6.5 and 8.0 but, above pH 8, inhibition by 2.5 mM Tris tends to diminish whereas inhibition by 40 mM PM increases abruptly to be essentially complete at pH 9.3 and above. As pH increases from 7.6 to 9.0, the apparent affinity of the free amine bases decreases whereas that of the cationic, protonated amines, increases. In this way, the protonated amines replace their corresponding free bases as the most potent inhibitors at high pH. The pH-dependent inhibition by 300 mM Li+ is essentially complete at pH 8, independent of the presence or absence of either 2.5 mM Tris or 40 mM PM. Even at pH 7.6, an excess (300 mM) of Li+ causes significant increases in the apparent Ki value of each Tris, PD (2-amino-2-methyl-1-3-propanediol) and PM, suggesting the possibility of a relation between the effects of Li+ and those of the hydroxylated amines which in fact are mutually exclusive inhibitors. The inhibitory results are interpreted in terms of a mechanistic model in which the free bases bind at two distinct sites in the enzyme's active center. Binding at the glucosyl sub-site occurs through the amine's free hydroxyl groups. This positioning facilitates the interaction between the lone electron pair of the deprotonated amino group with a proton donor in the enzyme's active center, characterized by a pK0 around 8.1. When this same group deprotonates, then the protonated amines acting as proton donors replace the free bases as the species giving fully competitive inhibition of sucrase.  相似文献   

14.
Treatment of submitochondrial particles (ETP) with trypsin at 0 degrees destroyed NADPH leads to NAD (or 3-acetylpyridine adenine dinucleotide, AcPyAD) transhydrogenase activity. NADH oxidase activity was unaffected; NADPH oxidase and NADH leads to AcPyAD transhydrogenase activities were diminished by less than 10%. When ETP was incubated with trypsin at 30 degrees, NADPH leads to NAD transhydrogenase activity was rapidly lost, NADPH oxidase activity was slowly destroyed, but NADH oxidase activity remained intact. The reduction pattern by NADPH, NADPH + NAD, and NADH of chromophores absorbing at 475 minus 510 nm (flavin and iron-sulfur centers) in complex I (NADH-ubiquinone reductase) or ETP treated with trypsin at 0 degrees also indicated specific destruction of transhydrogenase activity. The sensitivity of the NADPH leads to NAD transhydrogenase reaction to trypsin suggested the involvement of susceptible arginyl residues in the enzyme. Arginyl residues are considered to be positively charged binding sites for anionic substrates and ligands in many enzymes. Treatment of ETP with the specific arginine-binding reagent, butanedione, inhibited transhydrogenation from NADPH leads to NAD (or AcPyAD). It had no effect on NADH oxidation, and inhibited NADPH oxidation and NADH leads to AcPyAD transhydrogenation by only 10 to 15% even after 30 to 60 min incubation of ETP with butanedione. The inhibition of NADPH leads to NAD transhydrogenation was diminished considerably when butanedione was added to ETP in the presence of NAD or NADP. When both NAD and NADP were present, the butanedione effect was completely abolished, thus suggesting the possible presence of arginyl residues at the nucleotide binding site of the NADPH leads to NAD transhydrogenase enzyme. Under conditions that transhydrogenation from NADPH to NAD was completely inhibited by trypsin or butanedione, NADPH oxidation rate was larger than or equal to 220 nmol min-1 mg-1 ETP protein at pH 6.0 and 30 degrees. The above results establish that in the respiratory chain of beef-heart mitochondria NADH oxidation, NADPH oxidation, and NADPH leads to NAD transhydrogenation are independent reactions.  相似文献   

15.
In reaction centers from Rhodopseudomonas viridis that contain a single quinone, the decay of the photo-oxidized primary donor, P+, was found to be biphasic when the bound, donor cytochromes were chemically oxidized by ferricyanide. The ratio of the two phases was dependent on pH with an apparent pK of 7.6. A fast phase, which dominated at high pH (t1/2 = 1 ms at pH 9.5), corresponded to the expected charge recombination of P+ and the primary acceptor QA-. A much slower phase dominated at low pH and was shown to arise from a slow reduction of P+ by ferrocyanide in reaction centers where QA- has been rapidly oxidized by ferricyanide. The rate of QA- oxidation was linear with respect to ferricyanide activity and was strongly pH-dependent. The second-order rate constant, corrected for the activity coefficient of ferricyanide, approached a maximum of 2 X 10(8) M-1 X s-1 at low pH, but decreased steadily as the pH was raised above a pK of 5.8, indicating that a protonated state of the reaction center was involved. The slow reduction of P+ by ferrocyanide was also second-order, with a maximum rate constant at low pH of 8 X 10(5) M-1 X s-1 corrected for the activity coefficient of ferrocyanide. This rate also decreased at higher pH, with a pK of 7.4, indicating that ferrocyanide also was most reactive with a protonated form of the reaction center. The oxidation of QA- by ferricyanide was unaffected by the presence of o-phenanthroline, implying that access to QA- was not via the QB-binding site. In reaction centers supplemented with ubiquinone, oxidation of reduced secondary quinone, QB-, by ferricyanide was observed but was substantially slower than that for QA-. It is suggested that Q-B may be oxidized via QA so that the rate is modulated by the equilibrium constant for QA-QB in equilibrium with QAQB-.  相似文献   

16.
Bacterioferritin (EcBFR) of Escherichia coli is an iron-mineralizing hemoprotein composed of 24 identical subunits, each containing a dinuclear metal-binding site known as the "ferroxidase center." The chemistry of Fe(II) binding and oxidation and Fe(III) hydrolysis using H(2)O(2) as oxidant was studied by electrode oximetry, pH-stat, UV-visible spectrophotometry, and electron paramagnetic resonance spin trapping experiments. Absorption spectroscopy data demonstrate the oxidation of two Fe(II) per H(2)O(2) at the ferroxidase center, thus avoiding hydroxyl radical production via Fenton chemistry. The oxidation reaction with H(2)O(2) corresponds to [Fe(II)(2)-P](Z) + H(2)O(2) --> [Fe(III)(2)O-P](Z) + H(2)O, where [Fe(II)(2)-P](Z) represents a diferrous ferroxidase center complex of the protein P with net charge Z and [Fe(III)(2)O-P](Z) a micro-oxo-bridged diferric ferroxidase complex. The mineralization reaction is given by 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2FeOOH((core)) + 4H(+), where two Fe(II) are again oxidized by one H(2)O(2). Hydrogen peroxide is shown to be an intermediate product of dioxygen reduction when O(2) is used as the oxidant in both the ferroxidation and mineralization reactions. Most of the H(2)O(2) produced from O(2) is rapidly consumed in a subsequent ferroxidase reaction with Fe(II) to produce H(2)O. EPR spin trapping experiments show that the presence of EcBFR greatly attenuates the production of hydroxyl radical during Fe(II) oxidation by H(2)O(2), consistent with the ability of the bacterioferritin to facilitate the pairwise oxidation of Fe(II) by H(2)O(2), thus avoiding odd electron reduction products of oxygen and therefore oxidative damage to the protein and cellular components through oxygen radical chemistry.  相似文献   

17.
Timed assays in which GTP and GDP were separated and quantitated by HPLC were developed and used to study the metal activation of the mitochondrial and cytosolic isozymes of phosphoenolpyruvate carboxykinase purified from rabbit liver. These assays allowed both directions of catalysis to be studied under similar conditions and in the absence of coupling enzymes. The mitochondrial enzyme is rapidly inactivated by preincubation with Fe2+, as had been shown previously for the cytosolic isozyme. The greatest activation by Fe2+ was obtained by adding micromolar Fe2+ immediately after enzyme to form the complete assay mixture that also contained millimolar Mg2+. In the direction of synthesis of OAA from Pep, the K0.5 values for Mn2+ and Fe2+ were in the 3-7 microM range when a nonchelating buffer, Hepes, was used. The buffer used strongly affected activation by Fe2+ at pH 7.4; activation was eliminated in the case of phosphate and K0.5 increased several-fold over that obtained with Hepes when imidazole was used. In non-chelating buffer, the pH optimum was near 7.4 for both isozymes and for both directions of catalysis. However, the near optimal pH range extended below 7.4 for the direction of oxaloacetate synthesis while the range was above 7.4 for Pep synthesis. In the direction of oxaloacetate synthesis: (1) Both isozymes required the presence of micromolar Mn2+ or Fe2+ in addition to millimolar Mg2+ in order to shown significant activity. (2) Fe2+ was as effective an activator as Mn2+ at pH 7 and below. In the direction of Pep synthesis: (1) Micromolar Mn2+ was a much better activator than Fe2+ at the higher pH values needed for optimal activity in this direction. (2) With increasing pH, decreasing activation was obtained with Fe2+ while the activity supported by Mg2+ alone increased. The results demonstrate the potential for regulation of either isozyme of Pep carboxykinase by the availability of iron or manganese.  相似文献   

18.
Ferritin iron loading was studied in the presence of physiological serum phosphate concentrations (1 mM), elevated serum concentrations (2–5 mM), and intracellular phosphate concentrations (10 mM). Experiments compared iron loading into homopolymers of H and L ferritin with horse spleen ferritin. Prior to studying the reactions with ferritin, a series of control reactions were performed to study the solution chemistry of Fe2+ and phosphate. In the absence of ferritin, phosphate catalyzed Fe2+ oxidation and formed soluble polymeric Fe(III)-phosphate complexes. The Fe(III)-phosphate complexes were characterized by electron microscopy and atomic force microscopy, which revealed spherical nanoparticles with diameters of 10–20 nm. The soluble Fe(III)-phosphate complexes also formed as competing reactions during iron loading into ferritin. Elemental analysis on ferritin samples separated from the Fe(III)-phosphate complexes showed that as the phosphate concentration increased, the iron loading into horse ferritin decreased. The composition of the mineral that does form inside horse ferritin has a higher iron/phosphate ratio (~1:1) than ferritin purified from tissue (~10:1). Phosphate significantly inhibited iron loading into L ferritin, due to the lack of the ferroxidase center in this homopolymer. Spectrophotometric assays of iron loading into H ferritin showed identical iron loading curves in the presence of phosphate, indicating that the ferroxidase center of H ferritin efficiently competes with phosphate for the binding and oxidation of Fe2+. Additional studies demonstrated that H ferritin ferroxidase activity could be used to oxidize Fe2+ and facilitate the transfer of the Fe3+ into apo transferrin in the presence of phosphate.  相似文献   

19.
Cytochrome P450 (P450) 2D6 oxidizes a wide variety of drugs typically at a distance of 5-7 A from a basic nitrogen on the substrate. To investigate the determinants of P450 2D6 catalysis, we analyzed the binding and oxidation of phenethylamine substrates. P450 2D6 discriminated between the various phenethylamines, as evidenced by binding and steady-state results. Whereas the spectral binding affinity for 3-methoxyphenethylamine and 4-methoxyphenethylamine was similar, the affinity for 4-hydroxyphenethylamine was 12-fold weaker than for 3-hydroxyphenethylamine at pH 7.4. The binding of 3,4-dihydroxyphenethylamine was equally poor. These equilibrium dissociation constants were based on the observation of both type I and type II perturbation difference spectra; the former involves displacement of the proximal H(2)O ligand, yielding an iron spin state change, and the latter requires nitrogen ligation to the heme iron. One explanation for the observed type II binding spectra is the presence of both protonated and unprotonated forms of these compounds. To address this possibility, the K(S) values for 3-methoxyphenethylamine and 4-methoxyphenethylamine were determined as a function of pH. Two apparent pK(a) values were determined, which corresponded to a P450 2D6 residue involved in binding and to a lowered pK(a) of a substrate amine group upon binding P450 2D6. The apparent pK(a) of the enzyme residue (6.6) is much higher than the expected pK(a) of Asp301, which has been hypothesized to play a role in binding. Interestingly, the apparent pK(a) for the methoxyphenethylamine derivatives decreased by as much as 2 pH units upon binding to P450 2D6. 3-Methoxyphenethylamine and 4-methoxyphenethylamine underwent sequential oxidations with O-demethylation and subsequent ring hydroxylation to form 3,4-dihydroxyphenethylamine (dopamine). At higher substrate concentrations, the second oxidation was inhibited. This result can be explained by the increasing concentration of the inhibitory unprotonated substrate. Nevertheless, the rates of methoxyphenethylamine oxidations are the highest reported for P450 2D6 substrates.  相似文献   

20.
A comparative study of the rate of ferrocyanide-catalyzed oxidation of native sperm whale MbO2, its chemically modified derivative in which all accessible His residues are alkylated by sodium bromoacetate, (CM-MbO2), and mutant sperm whale MbO2 with His119 replaced by Asp residiue, [MbO2(His119-->Asp)] was carried out. The influence of pH, ionic strength, and [Fe(CN)6]4- concentration on the oxidation rate was investigated, as well as the effect of complexing MbO2 with redox-inactive Zn2+ ion, which, at the equimolar Zn2+ concentration, forms a stable complex with His119(GH1) on the protein surface. It was shown that the mechanism of the catalysis involves specific binding of [Fe(CN)6]4- to the protein at the His119(GH1) region, which is in agreement with a large positive electrostatic potential and the presence at this site of Mb of a cavity large enough to accommodate [Fe(CN)6]4- anion. The protonation of nearby His113 and His116 residiues (especially of the latter) plays a very important role in the catalysis, promoting the fast oxidation of bound [Fe(CN)6]4- by dissolved oxygen. Only the presence of these both necessary conditions in MbO2 structure provides its effective oxydation catalyzed by ferrocyanide.  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