首页 | 官方网站   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.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

2.
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   

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.
Stopped-flow kinetic studies of the formation of ferrioxamine B were performed. Formation of the complex follows the rate law
where Ka is the acid dissociation constant of the iron(III) aquo species in 0.1 M formate buffer. At 25°C k1 = 3.94 × 102M?1 sec?1, k2Ka = 1.18 × 10?1 sec?1, k3 = 3.6 × 10?1 sec?1. Activation parameters for k1 are ΔH = 11.7 kcal mole?1 and ΔS = ?8 cal K?1 mole?1. An associative mechanism is proposed. Attachment of the first chelate ring is the slow step and favorably positions the second chelate ring for attachment. Coordination of two chelate rings favorably positions the third chelate ring for attachment. These results are compared to kinetics of formation of model complexes and to a previous study of the formation of ferrioxamine B in which attachment of the third chelate ring was proposed as the slow step  相似文献   

5.
Genes of β-mannosidase 97 kDa, GH family 2 (bMann9), β-mannanase 48 kDa, GH family 5 (bMan2), and α-galactosidase 60 kDa, GH family 27 (aGal1) encoding galactomannan-degrading glycoside hydrolases of Myceliophthora thermophila C1 were successfully cloned, and the recombinant enzymes were purified to homogeneity and characterized. bMann9 displays only exo-mannosidase activity, the K m and k cat values are 0.4 mM and 15 sec?1 for p-nitrophenyl-β-D-mannopyranoside, and the optimal pH and temperature are 5.3 and 40°C, respectively. bMann2 is active towards galac-tomannans (GM) of various structures. The K m and k cat values are 1.3 mg/ml and 67 sec?1 for GM carob, and the optimal pH and temperature are 5.2 and 69°C, respectively. aGal1 is active towards p-nitrophenyl-α-D-galactopyranoside (PNPG) as well as GM of various structures. The K m and k cat values are 0.08 mM and 35 sec?1 for PNPG, and the optimal pH and temperature are 5.0 and 60°C, respectively.  相似文献   

6.
The stopped flow technique has been used to study the kinetics of complex formation of iron(III) with pyridoxal-5-phosphate (PLP) in the pH range 1.00–2.50, and in the temperature range 18 °C– 30 °C, at an ionic strength of. 0.50 M (NaCl). From the initial concentration dependence of PLP (TPLP,) of the reaction rate it can be shown that two kinetic steps can be represented as: kobs′ = mi + miPLP where mi and mi′ are pH-dependent parameters. The calculated activation data are δE* = 23.2 ± 1.8 kcal mol?1 and 10.98 ± 0.53 kcal mol?1 for the first and second kinetic steps, respectively and δS* are ?20.50 ± 5.96 e.u. and 24.62 ± 1.81 e.u., respeetively.  相似文献   

7.
Properties of [3H]diazepam binding sites on rat blood platelets   总被引:8,自引:0,他引:8  
J K Wang  T Taniguchi  S Spector 《Life sciences》1980,27(20):1881-1888
Intact rat blood platelets are shown to possess benzodiazepine binding sites of the peripheral type, binding of [3H]diazepam being strongly inhibited by Ro5-4864 (Ki = 3.6 ± 0.5 nM) but only weakly inhibited by clonazepam (Ki = 35.1 ± 18.2 μM). Binding of [3H]diazepam is specific and saturable. Scatchard analysis reveals a single class of binding sites with KD = 14.7 ± 1.0 nM and Bmax = 564 ± 75 fmoles/108 platelets. The Hill coefficient is 0.94, indicating a lack of binding site heterogeneity or negative cooperativity. Binding reaches equiliibrium at 6 min, with k+1 = 2.9 × 107 M?1 min?1, and is rapidly reversible (t12 = 2.2 min with K?1 = 0.315 min?1. KD derived from the rate constants agrees with that estimated by Scatchard analysis. KD of the crude membrane fraction of platelets is also close to that of intact platelets. Binding of [3H]diazepam is linear with platelet number (between 0.25–2 × 108 platelets), is temperature sensitive with maximum binding at 0°C, and has a broad optimal pH range between pH 5–9.  相似文献   

8.
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   

9.
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   

10.
The reactions of copper(II)-ahphatic 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 ammo acid oxidations, as changes in the rate law could be induced by modest variations in reductant concentration (at fixed [Cu(II)]o), pH, and the structure of the redox partners. With excess cysteine present at 25°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 × 103M?1sec?1), Cu(Me5dien)2+ (2.7 × 103M?1 sec?1), Cu(en)22+ (2.1 × 103M?1 sec?1), and Cu(dien)22+ (4.7 × 103 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 mtermediacy of di-μ-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°C) 7.5 × 107 M?2 sec?1, I = 0.2 M). 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 purposed, coupled with attack of a noncoordinated cysteine molecule on a vacated coordination position to stabilize the (Me6(tren)Cu(I) product.  相似文献   

11.
The interactions of chymotrypsin, subtilisin and trypsin with a low MW proteinase inhibitor from potatoes were investigated. The Ki value calculated for the binding of inhibitor to chymotrypsin was 1.6 ± 0.9 × 10?10M, while the second-order rate constant for association was 6 × 105 M?1/sec. Although binding was not observed to chymotrypsin which had been treated with diisopropyl fluorophosphate or with l-tosylamide-2-phenylethyl chloromethyl ketone, the 3-methylhistidine-57 derivative bound inhibitor with a Ki value of 9.6 × 10?9 M. The inhibitor also exhibited a tight association with subtilisin (Ki < 4 × 10?9 M). In contrast, little inhibition of trypsin was observed, and this was believed to be due to low levels of a contaminant in our preparations. No evidence for reactive site cleavage was observed after incubation of the inhibitor with catalytic amounts of chymotrypsin or subtilisin at acid pH.  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.
The translational diffusion coefficient of CF1 at low and high protein concentration as well as at different ionic strength (0.05 – 1.65 M) wsa determined by means of quasi-elastic light scattering experiments. The diffusion coefficient changes from D20,wo = 3.12 × 10?7 cm2 · sec?1 at 0.05 M, pH 7.8, 20°C, to D20,wo = 3.52 × 10?7 cm2 · sec?1 at 1.6 M, pH 7.8, 20°C. At high enzyme concentration (20 mg/ml) and under crystallization conditions (Paradies, BBRC 91: 685, 1979) CF1 behaves as a solution of “true” hard spheres, whereas at low salt concentration the ionic atmosphere has a larger spatial extent, resulting in a higher effective hydrodynamic radius (RH = 65 Å).  相似文献   

14.
The anation reaction of aquopentaamminecobalt(III) by acetate has been studied in the temperature range 60–80°C and acidity range 1.0 ≦ pH ≦ 5.5 for total acetate concentrations ≦ 0.5 M and at ionic strength 1.0 M. The anation by acetic acid follows second-order kinetics (k0), whereas the kinetic results for the anation by acetate (Q1, k1) provide evidence for the formation of an ion-pair with the complex ion. Typical experimental results at 70°C are k0 = 5.33 X 10−5 M−1 sec−1, Q1 = 5.87 M−1 and k1 = 1.46 X 10−4sec−1. The activation parameters for the different reaction paths are reported and the results discussed with reference to various other anation reactions of Co(III) complexes.  相似文献   

15.
The reaction of parsley 2Fe-2S ferredoxin in the normal oxidized state with eaq? generated by pulse radiolysis techniques has been studied at ~25°C, pH 7–8, I = 0.10 M (NaClO4). Rate constants ke (eaq? decay) and kp (protein absorbance change) are the same, second-order rate constant 9.7 × 109 M?1 sec?1. The reaction exhibits close to 100% efficiency. With 8Fe-8S ferredoxin from Clostridium pasteurianum under identical conditions it now appears that kp (although sometimes significantly smaller) is equal to ke. Varying efficiencies are also observed with this protein depending on the batch used. The reasons for such variable behavior are not fully understood. With oxidized and reduced forms of Chromatium v. high-potential iron-sulfur protein (HIPIP), ke and kp are essentially the same, but the highest efficiency observed is only ~50%. The prevailing pattern is therefore that rate constants ke and kp are generally in step for proteins having a single (or identical) active site(s). When the active site is buried as with HIPIP the efficiency of the reaction appears to decrease.  相似文献   

16.
Studies of the binding of Ni2+ to adenylyl-3',5'-adenosine (ApA) at pH 6-0 by ultraviolet spectrophotometry indicate the formation of a 1:1 complex in the presence of a large excess of metal ion. At 25 °C. and ionic strength μ = 0.5 M, the stability constant of Ni(ApA) is evaluated to be K = 2.6 (±0.6) M?1. The low stability is taken as evidence that the predominant complex species is one in which the ApA acts as a monodentate ligand, mainly through the adenine group. The rate constants for complex formation and dissociation, kf = 1430 M?1 s?1 and kb = 665 s?1 (25°C. μ = 0.5M). determined by the temperature-jump relaxation technique, are consistent with this interpretation. The binding strength of Ni2+ to poly(adenylic acid) [poly(A)] has been studied at pH 7.0 using murexide as an indicator of the concentration of free Ni2+. Within the concentration range [Ni2+ = 1 × 10?5 × 10?3 M the data can be represented in the form of a linear Scatchard plot. i.e., the process can be described as the binding of Ni2+ to one class of independent binding sites. The number of binding sites per monomer is 0.26, and the stability constant K = 8.2×103 M?1 (25°C μ = 0.1 M). In kinetic studies of the reaction of Ni2+ with poly(A), two relaxation effects due to complex formation were detected, one with a concentration-independent time constant of about 0.4 ms, the other with a concentration-dependent time constant in the millisecond range. The concentration dependence of the longer relaxation time can be accounted for by a three-step mechanism which consists of a fast second-order association reaction followed by two first-order steps. There is evidence, however, that the overall process is more complicated than expressed by the three-step mechanism.  相似文献   

17.
Association of a sulfated galactosyl ceramide, sulfatide, with the viral envelope glycoprotein hemagglutinin (HA) delivered to the cell surface is required for influenza A virus (IAV) replication through efficient translocation of the newly synthesized viral nucleoprotein from the nucleus to the cytoplasm. To determine whether the ectodomain of HA can bind to sulfatide, a secreted-type HA (sHA), in which the transmembrane region and cytoplasmic tail were deleted, was generated by using a baculovirus expression system. The receptor binding ability and antigenic structure of sHA were evaluated by a hemagglutination assay, solid-phase binding assay and hemagglutination inhibition assay. sHA showed subtype-specific antigenicity and binding ability to both sulfatide and gangliosides. Kinetics of sHA binding to sulfatide and GD1a was demonstrated by quartz crystal microbalance (QCM) analysis. QCM analysis showed that the sHA bound with the association rate constant (k on) of 1.41?×?104 M?1 sec?1, dissociation rate constant (k off) of 2.03?×?10?4 sec?1 and K d of 1.44?×?10?8 M to sulfatide immobilized on a sensor chip. The k off values of sHA were similar for sulfatide and GD1a, whereas the k on value of sHA binding to sulfatide was 2.56-times lower than that of sHA binding to GD1a. The results indicate that sulfatide directly binds to the ectodomain of HA with high affinity.  相似文献   

18.
Quaterpyridyneiron (III) complex ions anchored to partially ordered poly (L-glutamate) or poly (D-glutamate) were used as (enantiomeric) catalysts for the H2O2-oxidation of L(+) ascorbic acid at pH 7. When the α-helical fraction of polypeptide matrices was low, the configuration dissymmetry of the active sites was unable to impart any stereoselective effect in the catalysis, i.e. k = 3.66 x 103 M?1?sec?1 (25.9°C) with both catalysts. On the contrary, by increasing the amount of α-helix in the polymeric supports the stereoselectivity increases, the second-order rate constants kFeD being definitely higher than kFeL.Implications of the role played by the conformational dissymmetry of the active sites in the stereospecificity of the process are briefly discussed.  相似文献   

19.
The nonenzymatic reduction of nitrosobenzene by NADPH and NADH in aqueous buffer solution at 25°C is described. Both reactants quantitatively convert nitrosobenzene to phenylhydroxylamine. Rate constants for reduction (kr) were determined spectrophotometrically and found to be identical at pH 5.7 and 7.4 and independent of buffer concentration. The values of kNADH (124–149 M?1 sec?1) and kNADPH (131–170 M?1 sec?1) are essentially identical. The reaction is not subject to general catalysis or specific salt effects. The oxidation of phenylhydroxylamine by NAD(P) to nitrosobenzene is only stimulated by a factor of 1.2 over oxidation in its absence (when the ratio of NADP: phenylhydroxylamine was 8:1).  相似文献   

20.
β-Xylosidase was purified 662 fold from a culture filtrate by ammonium sulfate fractionation, gel filtration on Biogel P-100, DEAE-Sephadex chromatography, and gel filtration on Sephadex G-200. With isoelectric focusing, the purified β-xylosidase found to be homogeneous on SDS (sodium dodecyl sulfate) polyacrylamide gel electrophoresis. The molecular weight was estimated by gel filtration to be 240,000, and 116,000 by SDS polyacrylamide gel electrophoresis. The purified β-xylosidase had an isoelectric point at pH 3.25, and contained 4% carbohydrate residue. The optimum pH was found to be in the range of 4.5 ~ 5, and the optimum temperature was 55°C. The enzyme activity was inhibited by Hg2 +, SDS, and N-bromosuccinimide at a concentration of 1 × 10?3 m, and also p-chloromercuribenzoate at a concentration of 1 × 10?4m. The purified enzyme hydrolyzed phenyl β-d-xyloside (ko = 302.6 sec?1),β-nitrophenyl β-d-xyloside (ko = 438.9 sec?1), o-nitrophenyl β-d-xyloside (ko = 431.0 sec?1), p-chlorophenyl β-d-xyloside (ko = 207.9 sec?1), o-chlorophenyl β-d-xyloside (ko = 211.8 sec?1), β-methylphenyl β-d-xyloside ko = 96.5 sec?1), o-methylphenyl β-d-xyloside (ko = 83.1 sec?1), p-methoxyphenyl β-d-xyloside (ko = 99.3 sec?1), o-methoxyphenyl β-d-xyloside (ko= 100.0 sec?1), xylobiose (ko = 992A sec?1), xylotriose (ko = 1321.9 sec?1), xylotetraose (ko = 7S9.1 sec?1) and xylopentaose (ko = 508.0 sec?1). On enzymic hydrolysis of phenyl β-d-xyloside, the reaction product was found to be β-d-xylose with retention of the configuration. The purified β-xylosidase was practically free of a-xylosidase and β-glucosidase activities.  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-23

京公网安备 11010802026262号