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 656 毫秒
1.
The reactions of copper(II)-aliphatic polyamine complexes with cysteine, cysteine methyl ester, penicillamine, and glutathione have been investigated, with the goal of understanding the relationship between RS- -Cu(II) adduct structure and preferred redox decay pathway. Considerable mechanistic flexibility exists within this class of mercapto amino acid oxidations, as changes in the rate law could be induced by modest variations in reductant concentration (at fixed [Cu(II)]0), pH, and the structure of the redox partners. With excess cysteine present at 25 degrees C, pH 5.0, I = 0.2 M (NaOAc), decay of 1:1 cys-S- -Cu(II) transient adducts was found to be first order in both cys-SH and transient. Second-order rate constants characteristic of Cu(dien)2+(6.1 X 10(3) M-1 sec-1), Cu(Me5dien)2+ (2.7 X 10(3) M-1 sec-1), Cu(en)22+ (2.1 X 10(3) M-1 sec-1), and Cu(dien)22+ (4.7 X 10(3) M-1 sec-1) are remarkably similar, considering substantial differences in the composition and geometry of the oxidant first coordination sphere. A mechanism involving attack of cysteine on the coordinated sulfur atom of the transient, giving a disulfide anion radical intermediate, is proposed to account for these results. Moderate reactivity decreases in the cysteine-Cu(dien)2+, Cu(Me5dien)2+ reactions with increasing [H+] (pH 4-6) reflect partial protonation of the polyamine ligands. A very different rate law, second order in the RS- -Cu(II) transient and approximately zeroth order in mercaptan, applies in the pH 5.0 oxidations of cysteine methyl ester, penicillamine, and glutathione by Cu(dien)2+ and Cu(Me5dien)2+. This behavior suggests the intermediacy of di-mu-mercapto-bridged binuclear Cu(II) species, in which a concerted two-electron change yields the disulfide and Cu(I) products. Similar hydroxo-bridged intermediates are proposed to account for the transition from first- to second-order transient dependence in cysteine oxidations by Cu(dien)2+ and Cu(Me5dien)2+ as the pH is increased from 5 to 7. Yet another rate law, second order in transient and first order in cysteine, applies in the pH 5.0 oxidation of cysteine by Cu(Me6tren)2+ (k(25 degrees C) 7.5 X 10(7) M-2 sec-1, I = 0.2M). Steric rigidity of this trigonal bipyramidal oxidant evidently protects the coordinated sulfur atom from attack in a RSSR- -forming pathway. Formation of a coordinated disulfide in the rate-determining step is proposed, coupled with attack of a noncoordinated cysteine molecule on a vacated coordination position to stabilize the (Me6tren)Cu(I) product.  相似文献   

2.
The rate constants of the reactions between pulse radiolytically produced superoxide anions and the Cu(II) chelates of salicylate, acetylsalicylate, p-aminosalicylate and diisopropylsalicylate were determined at pH 7.5 and found to range from 0.8 to 2.4 × 109 M?1 sec?1. It was intriguing to note that they had a superoxide dismutase activity identical with that of native cuprein-copper (k245 = 1.3 × 109 M?1 sec?1 per g-atom of Cu). These measurements confirm our earlier observations using indirect assays that all copper salicylates act as perfect model superoxide dismutases and favour the proposal that the activity of anti-inflammatory agents might be assigned to their in vivo formed Cu complexes.  相似文献   

3.
Kinetic studies of the reduction of ferrioxamine B (Fe(Hdesf)+) by Cr(H2O)62+, V(H2O)62+, and dithionite have been performed. For Cr(H2O)62+ and V(H2O)62+, the rate is ?d[Fe(Hdesf)+]/dt = k[Fe(Hdesf)+][M2+]. For Cr(H2O)62+, k = 1.19 × 104 M?1 sec?1 at 25°C and μ = 0.4 M, and k is independent of pH from 2.6 to 3.5. For V(H2O)62+, k = 6.30 × 102 M?1 sec?1 at 25°C, μ = 1.0 M, and pH = 2.2. The rate is nearly independent of pH from 2.2 to 4.0. For Cr(H2O)62+ and V(H2O)62+, the activation parameters are ΔH = 8.2 kcal mol?1, ΔS ?12 eu and ΔH = 1.7 kcal mol?1, ΔS = ?40 eu (at pH 2.2) respectively. Reduction by Cr(H2O)62+ is inner-sphere, while reduction by V(H2O)62+ is outer-sphere. Reduction by dithionite follows the rate law ?d[Fe(Hdesf)+]/dt =kK12[Fe(Hdesf)+][S2O42?]12 where K is the equilibrium constant for dissociation of S2O42? into SO2? radicals. The value of k at 25°C and μ = 0.5 is 2.7 × 103 M?1 sec?1 at pH 5.8, 3.5 × 103 M?1 sec?1 at pH 6.8, and 4.6 × 103 M?1 sec?1 at pH 7.8, and ΔH = 6.8 kcal mol?1 and ΔS = ?19 eu at pH 7.8.  相似文献   

4.
The electron transfer reactions of horse heart cytochrome c with a series of amino acid-pentacyanoferrate(II) complexes have been studied by the stopped-flow technique, at 25°C, μ = 0.100, pH 7 (phosphate buffer). A second-order behavior was observed in the case of the Fe(CN)5 (histidine)3? complex, with k = 2.8 x 105 M?1 sec?1. For the Fe(CN)5 (alanine)4? and Fe(CN)5(L-glutamate)5? complexes, only a minor deviation of the second-order behavior, close to the experimental error (k = 3.2 × 105 and 1.6 x 105 M?1 sec?1, respectively) was noted at high concentrations of the reactants (e.g., 6 × 10?4 M). The results are in accord with recent work on the Fe(CN)64?/cytochrome c system demonstrating weak association of the reactants. The calculated self-exchange rate constants including electrostatic interactions for the imidazole,L -histidine, 4-aminopyridine, glycinate, β-alaninate, andL-glutamate pentacyanoferrate(II) complexes were 3.3 × 105, 3.3 × 105, 2.8 × 106,4.1 × 102,5.5 × 102, and 6.0 M?1 sec?1, respectively. Marcus theory calculations for the cytochrome c reactions were interpreted in terms of two nonequivalent binding sites for the complexes, with the metalloprotein self-exchange rate constants varying from 104 M?1 sec?1 (histidine, imidazole, and 4-aminopyridine complexes) to 106 M?1 sec ?1 (glycinate, β-alaninate, and L-glutamate complexes).  相似文献   

5.
A kinetic study of the oxidation of (hydroxyethyl)ferrocene (HEF) by [2-pyridylmethylbis(2-ethyl-thioethyl)ainine]copper(II) (Cu(pmas)2+) is reported, with the objective of documenting the influence of the two thioether sulfur ligands on the electron transfer rate. Both reactants exhibit a first-order dependence at pH 6, I = 0.1 M(NaNO3); k(25°C) = 1.3 × 104M−1sec−1, ΔH3 = 10.1 kcal/mole, ΔS3 = −6 eu. The apparent Cu(pmas)2+/+ self-exchange electron transfer rate constant calculated from this reaction on the basis of relative Marcus theory (4.7 × 101M−1 sec−1) agrees well with previous findings on ferrocytochrome c, Fe(CN)64−, and Ru(NH3)5py2+ oxidations. Spectrophotometric titrations of Cu(pmas)2+ and Cu(tmpa)2+ (tmpa = tris(2-pyridylmethyl)amine) with azide ion showed that both Cu(pmas)N3)+ (Kf1 = 3.1 × 103M−1) and Cu(pmas)(N3)2 (Kf2 = 3.5 × 101M−1) but Cu(tmpa)(N3)+ (Kf = 6.6 × 102M−1) are formed up to 0.15 M N3 (25°C, pH 6, I = 0.2 M), suggesting that a thioether sulfur atom is displaced in the uptake of a second N3 ion by Cu(pmas)(N3)+. The effect of thioether sulfur displacement by azide ion on the HEF-Cu(pmas)2+ reaction rate may be understood entirely through the tendency of N3 to shift the position of the redox equilibrium towards the reactant side, without invoking any special role for the sulfur ligand in promoting electron transfer reactivity.  相似文献   

6.
The stoichiometry and rate of oxidation with dioxygen of tetra-(p-sulfonatophenyl)-porphinatomanganese(II) and the bisimidazole tetra(p-sulfonatophenyl)porphinato-iron(II) were studied in aqueous solutions at neutral pH. The stoichiometry for both complexes was determined; two molecules of metalloporphyrin reacted with dioxygen to produce the +3 oxidation state of the metalloporphyrins and hydrogen peroxide. The rate law for the oxidation of Mn(II)-TPPS is rate = k′[Mn(II)-TPPS][O2], with k′ at 26.5° of 2.6 × 105 M?1 sec?1. The rate law for the oxidation of Fe(II)-TPPS in the presence of imidazole is
with k″ = 10,100 sec?1. Some possible mechanisms consistent with these data are discussed.  相似文献   

7.
《Inorganica chimica acta》1988,149(1):139-145
The stoichiometry and kinetics of the reaction between [Cu(dien)(OH)]+ and [Fe(CN)6]3− in aqueous alkaline medium are described. The rate equation − (d[Fe(III)]/dt = {k1[OH]2[[Cu(dien)(OH)]+] + k2[OH] × [[Cu(dien)(OH)]+]2}([Fe(III)]/[Fe(II)]) (Fe(III) = [Fe(CN)6]3−; Fe(II) = [Fe(CN)6]4−, the 4:4:1 OH/Fe(III)/[Cu(dien)(OH)]+ stoichiometric ratio and the nature of the ultimate products identified in the reaction solution suggest the fast formation of a doubly deprotonated Cu(III)-diamido complex which slowly undergoes an internal redox process where the ligand is oxidised to the Schiff base H2NCH2CH2NCHCHNH.The [[Cu(dien)(OH)]+]2 term in the rate equation is explained with the formation of a transient μ-hydroxo mixed-valence Cu dimer. A two-electron internal reduction of the Cu(III) complex yielding a Cu(I) intermediate is suggested to account for the presence of monovalent copper in a precipitate which forms at relatively high reactant concentrations and in the absence of dioxygen.  相似文献   

8.
The hydrolyses of p-nitrotrifluoroacetanilide catalyzed by water and imidazole were examined at 70°C. The pH-rate constant profile of the hydrolysis in H2O was examined in the pH range 0.0–11.4. The hydrolysis was independent of pH in the region from pH 1.0 to 4.5, presumably a water-catalyzed reaction. The rate constant and the D2O solvent isotope effect for this reaction were 1.0 × 10?4 sec?1 and 3.7, respectively. Both natural imidazole and imidazolium cation catalyzed hydrolysis. The rate constant of the hydrolysis catalyzed by neutral imidazole was determined to be 5.4 × 10?3M?1 sec?1 and the D2O solvent isotope effect was 1.8.  相似文献   

9.
Oxymyoglobin reacts with imidazole, substituted imidazoles, and hydroquinone to give metmyoglobin. The kinetics of these reactions have been studied. The rates are first order in both reactants, and second-order rate constants are reported. At pH 8.2, k1 for imidazole is 2.5 ± 0.3 × 10?3 M?1 sec?1 and for hydroquinone is 4 ± 0.4 × 10?1 M?1 sec?1. The rates are independent of pH for imidazole but increase rapidly with pH for hydroquinone. The mechanism for all these reactions is thought to involve the two-electron reduction of molecular oxygen to peroxide with concurrent oxidation of both the protein and the reactant. An analogous mechanism has been suggested previously [1] for the reaction of oxyhemoglobin with hydroquinone. It has previously been shown [6] that imidazole can mediate the transfer of electrons to heme proteins by forming a transient reduced radical. The present results indicate that it can also form a transient oxidized radical under mild conditions. This dual capability may be important in biological electron-transfer processes.  相似文献   

10.
A bulky platinum triamine complex, [Pt(Me5dien)(NO3)]NO3 (Me5dien = N,N,N′,N′,N′′-pentamethyldiethylenetriamine) has been prepared and reacted in D2O with N-acetylmethionine (N-AcMet) and guanosine 5′-monophosphate (5′-GMP); the reactions have been studied using 1H NMR spectroscopy. Reaction with 5′-GMP leads to two rotamers of [Pt(Me5dien)(5′-GMP-N7)]+. Reaction with N-AcMet leads to formation of [Pt(Me5dien)(N-AcMet-S)]+. When a sample with equimolar mixtures of [Pt(Me5dien)(D2O)]2+, 5′-GMP, and N-AcMet was prepared, [Pt(Me5dien)(5′-GMP-N7)]+ was the dominant product observed throughout the reaction. This selectivity is the opposite of that observed for a similar reaction of [Pt(dien)(D2O)]2+ with 5′-GMP and N-AcMet. To our knowledge, this is the first report of a platinum(II) triamine complex that reacts substantially faster with 5′-GMP than with N-AcMet; the effect is most likely due to steric clashes between the methyl groups of the Me5dien ligand and the N-AcMet.  相似文献   

11.
The acid-catalyzed hydrolysis of heparin from Cu(II) complex was studied as a function of time and temperature. Four independent calculations showed that the hydrolysis, during the 5-hr period examined, obeys the first-order kinetic law. Specific rate constants, calculated at 50°C, 57°C, 65°C, 71°C, and 80°C, were 3.3 × 10?5 sec?1, 6.5 × 10?5 sec?1, 10.4 × 10?5 sec?1, 15.1 × 10?5 sec?1, and 26.6 × 10?5 sec?1, respectively. Arrhenius plots of the data yielded 14.7 kcal as the energy of activation. An independent run of the self-hydrolysis of heparin at 57°C also obeyed first-order kinetics and its specific rate constant of 6.4 × 10?5 sec?1 is in excellent agreement with that of the hydrolysis of Cu(II)-heparin at 57°C. The anticoagulant activity of heparin and of the Cu(II)-heparin are not appreciably different. Further, the inactivation of heparin closely parallels Cu(II) release from the Cu(II) complex which in turn parallels desulfation.  相似文献   

12.
The kinetics of the reduction by aniline and a series of substituted anilines of a peroxidatically active intermediate, formed by oxidation of deuteroferriheme with hydrogen peroxide, have been studied by stopped-flow spectrophotometry. The reaction with aniline was first order with respect to [intermediate] and showed first-order saturation kinetics with respect to [aniline]. The second-order rate constant was 2.0 ± 0.2 × 105 M?1 sec?1 at 25°C (independent of pH in the range 6.60–9.68) compared with the value of 2.4 × 105 M?1 sec?1 for the reaction of aniline with horseradish peroxidase Compound I. The effect of aniline substituents upon reactivity towards the heme intermediate closely paralled those reported for reaction with the enzymic intermediate. Anilines bearing electron-donating substituents reacted more rapidly and those bearing electron-withdrawing substituents more slowly than the unsubstituted amine. The rate constants for the heme intermediate reactions (kdfh)found to be related to those for the enzymic reactions (khrp) by the equation:log kDFH= 0.65log kHRP+ 1.96 with a correlation coefficient of 0. 98.  相似文献   

13.
Two trypsin inhibitors, LA-1 and LA-2, have been isolated from ridged gourd (Luffa acutangula Linn.) seeds and purified to homogeneity by gel filtration followed by ion-exchange chromatography. The isoelectric point is atpH 4.55 for LA-1 and atpH 5.85 for LA-2. The Stokes radius of each inhibitor is 11.4 å. The fluorescence emission spectrum of each inhibitor is similar to that of the free tyrosine. The biomolecular rate constant of acrylamide quenching is 1.0×109 M?1 sec?1 for LA-1 and 0.8 × 109 M?1 sec?1 for LA-2 and that of K2HPO4 quenching is 1.6×1011 M?1 sec?1 for LA-1 and 1.2×1011M?1 sec?1 for LA-2. Analysis of the circular dichroic spectra yields 40%α-helix and 60%Β-turn for La-1 and 45%α-helix and 55%Β-turn for LA-2. Inhibitors LA-1 and LA-2 consist of 28 and 29 amino acid residues, respectively. They lack threonine, alanine, valine, and tryptophan. Both inhibitors strongly inhibit trypsin by forming enzymeinhibitor complexes at a molar ratio of unity. A chemical modification study suggests the involvement of arginine of LA-1 and lysine of LA-2 in their reactive sites. The inhibitors are very similar in their amino acid sequences, and show sequence homology with other squash family inhibitors.  相似文献   

14.
The kinetics of formation of the intermediate complex between catalase and H2O2 has been reexamined. It has been shown that the kinetics consists of a rapid and of a subsequent slow phase. At the maximum of the transient decrement of the optical absorption, the system was found to be in a terminal state with regard to the rapid phase. On this basis, the formation curve of the intermediate complex was calculated. From the parameters of the curve the maximal saturation of catalase hematins (from horse erythrocytes) by H2O2 is 35%. The absolute spectrum of the intermediate complex was established. The variation of the previously calculated rate constant of formation of the intermediate complex was shown to be due to the inapplicability of the pre-steady-state approximation to the rate data. By applying a more general approach and by the use of a computer, the individual rate constants of the peroxidatic scheme were calculated (relevant to micromolar solutions of catalase) k1 = (3.0 ± 0.2) × 106 M?1 sec?1k4 = (5.6 ± 0.3) × 106 M?1 sec?1 These values are 2.2 times higher in a nanomolar solution.  相似文献   

15.
Type II extradiol dioxygenase, 2′-carboxy-2,3-dihydroxybiphenyl 1,2-dioxygenase (FlnD1D2) involved in the fluorene degradation pathway of Rhodococcus sp. DFA3 was purified to homogeneity from a heterologously expressing Escherichia coli. Gel filtration chromatography and SDS-PAGE suggested that FlnD1D2 is an α4β4 heterooctamer and that the molecular masses of these subunits are 30 and 9.9 kDa, respectively. The optimum pH and temperature for enzyme activity were 8.0 and 30 °C, respectively. Assessment of metal ion effects suggested that exogenously supplied Fe2+ increases enzyme activity 3.2-fold. FlnD1D2 catalyzed meta-cleavage of 2′-carboxy-2,3-dihydroxybiphenyl homologous compounds, but not single-ring catecholic compounds. The Km and kcat/Km values of FlnD1D2 for 2,3-dihidroxybiphenyl were 97.2 μM and 1.5 × 10?2 μM?1sec?1, and for 2,2′,3-trihydroxybiphenyl, they were 168.0 μM and 0.5 × 10?2 μM?1sec?1, respectively. A phylogenetic tree of the large and small subunits of type II extradiol dioxygenases suggested that FlnD1D2 constitutes a novel subgroup among heterooligomeric type II extradiol dioxygenases.  相似文献   

16.
《Free radical research》2013,47(11):1300-1310
Abstract

Hypotaurine and cysteine sulfinic acid are known to be readily oxidized to the respective sulfonates, taurine and cysteic acid, by several oxidative agents that may be present in biological systems. In this work, the relevance of both the carbonate anion and nitrogen dioxide radicals in the oxidation of hypotaurine and cysteine sulfinic acid has been explored by the peroxidase activity of Cu,Zn superoxide dismutase (SOD) and by pulse radiolysis. The extent of sulfinate oxidation induced by the system SOD/H2O2 in the presence of bicarbonate (CO3?– generation), or nitrite (?NO2 generation) has been evaluated. Hypotaurine is efficiently oxidized by the carbonate radical anion generated by the peroxidase activity of Cu,Zn SOD. Pulse radiolysis studies have shown that the carbonate radical anion reacts with hypotaurine more rapidly (k = 1.1 × 109 M?1s?1) than nitrogen dioxide (k = 1.6 × 107 M?1s?1). Regarding cysteine sulfinic acid, it is less reactive with the carbonate radical anion (k = 5.5 × 107 M?1s?1) than hypotaurine. It has also been observed that the one-electron transfer oxidation of both sulfinates by the radicals is accompanied by the generation of transient sulfonyl radicals (RSO2?). Considering that the carbonate radical anion could be formed in vivo at high level from bicarbonate, this radical can be included in the oxidants capable of performing the last metabolic step of taurine biosynthesis. Moreover, the protective effect exerted by hypotaurine and cysteine sulfinate on the carbonate radical anion-mediated tyrosine dimerization indicates that both sulfinates have scavenging activity towards the carbonate radical anion. However, the formation of transient reactive intermediates during sulfinate oxidation by carbonate anion and nitrogen dioxide radical may at the same time promote oxidative reactions.  相似文献   

17.
The complexation reactions of O-phospho-DL-serine with Ni(II) or Co(II) were studied at 25°C and ionic strength 0.2 M (KNO3) by temperature-jump. The observed rate constants for formation of the Ni2+ and Co+2 monocomplexes were (1.32 ± 0.09) × 105 and (1.73 ± 0.33) × 107 M?1 sec?1, respectively. Complexation is postulated to involve formation of a monocoordinated steady state intermediate followed by rate-determining chelate ring closure.  相似文献   

18.
A two-dimensional copper(II) polymer with formula of [Cu4(H2O)4(dmapox)2(btc)]n · 10nH2O, where dmapox is the dianion of N,N′-bis[3-(dimethylamino)propyl]oxamide and btc is the tetra-anion of 1,2,4,5-benzenetetracarboxylic acid, was synthesized and characterized by elemental analysis, conductivity measurement, IR and electronic spectral studies. The crystal structure of the complex has been determined by X-ray single-crystal diffraction. The structure consists of crystallized water molecules and neutral two-dimensional copper(II) coordination polymeric networks constructed both by the bis-tridentate μ-trans-dmapox and tetra-monodentate μ4-btc bridging ligands. Each btc ligand links four trans-dmapox-bridged binuclear copper(II) building blocks [Cu2(H2O)2(trans-dmapox)]2+ and each binuclear copper(II) building block attaches to two btc ligands forming an infinite 2D layer which consists of 4+4 grids with dimensions of 13.563(5) × 15.616(5) Å. The environment around the copper(II) atom can be described as a distorted square-pyramid and the Cu?Cu separations through μ-trans-dmapox and μ4-btc bridging ligands are 5.225 Å (Cu1-Cu1i), 5.270 Å (Cu2-Cu2ii), 6.115 Å (Cu1-Cu2), 9.047 Å (Cu1-Cu2iii) and 10.968 Å (Cu1-Cu1iii), respectively. Abundant hydrogen bonds among the crystallized, the coordinated water molecules, and the uncoordinated carboxyl oxygen atoms cross-link the two-dimensional layers into an overall three-dimensional channel-like framework. The interaction of the copper(II) polymer with calf thymus DNA (CT-DNA) has been investigated by using absorption, emission spectral and electrochemical techniques. The results indicate that the copper(II) polymer interacts with DNA strongly (Kb = 4.8 × 105 M−1 and Ksv = 1.1 × 104) and the interaction mode between the copper(II) polymer and DNA may be the groove binding. To the best of our knowledge, this is the first report about the crystal structure and DNA-binding studies of a two-dimensional copper(II) polymer bridged both by the trans-oxamidate and btc ligands.  相似文献   

19.
The ferric hemes of valence hybrid hemoglobins combine with imidazole in a manner analogous with the hemes of methemoglobin. Equilibrium studies show that imidazole binding to methemoglobin is minimally described by the sum of two independent processes (K1 = 200 M?1 and K2 = 37 M?1), both of which contribute equally to the observed difference spectrum. Using valance hybrid hemoglobins, which show single binding processes under similar conditions, it is possible to identify the high affinity sites in methemoglobin with the α chains and the low affinity sites with the β chains.Kinetic studies show that the valance hybrid hemoglobins react in a single exponential fashion with imidazole in contrast with methemoglobin which shows a biphasic reaction (k1 = 85 M?1 sec?1k2 = 25 M?1 sec?1). A comparison of the rates of reaction of the hybrids allows the assignment of the fast phase in methemoglobin to the β chains and the slow phase to the α chains.The heterogeneity of the imidazole reaction with methemoglobin occurs over the pH range 5.5–9.5 within which two ionization processes are discernable at pH 6.9 and 7.5.  相似文献   

20.
Horse heart carboxymethylated cytc (CM-cytc) displays myoglobin-like properties. Here, the effect of cardiolipin (CL) liposomes on the nitrite reductase activity of ferrous CM-cytc [CM-cytc-Fe(II)], in the presence of sodium dithionite, is reported between pH 5.5 and 7.6, at 20.0 °C. Cytc-Fe(II) displays a very low value of the apparent second-order rate constant for the NO2 ?-mediated conversion of cytc-Fe(II) to cytc-Fe(II)-NO [k on = (7.3 ± 0.7) × 10?2 M?1 s?1; at pH 7.4], whereas the value of k on for NO2 ? reduction by CM-cytc-Fe(II) is 1.1 ± 0.2 M?1 s?1 (at pH 7.4). CL facilitates the NO2 ?-mediated nitrosylation of CM-cytc-Fe(II) in a dose-dependent manner, the value of k on for the NO2 ?-mediated conversion of CL–CM-cytc-Fe(II) to CL–CM-cytc-Fe(II)-NO (5.6 ± 0.6 M?1 s?1; at pH 7.4) being slightly higher than that for the NO2 ?-mediated conversion of CL–cytc-Fe(II) to CL–cytc-Fe(II)-NO (2.6 ± 0.3 M?1 s?1; at pH 7.4). The apparent affinity of CL for CM-cytc-Fe(II) is essentially pH independent, the average value of B being (1.3 ± 0.3) × 10?6 M. In the absence and presence of CL liposomes, the nitrite reductase activity of CM-cytc-Fe(II) increases linearly on lowering pH and the values of the slope of the linear fittings of Log k on versus pH are ?1.05 ± 0.07 and ?1.03 ± 0.03, respectively, reflecting the involvement of one proton for the formation of the transient ferric form, NO, and OH?. These results indicate that Met80 carboxymethylation and CL binding cooperate in the stabilization of the highly reactive heme-Fe atom of CL–CM-cytc.  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