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.
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   

2.
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   

3.
《Inorganica chimica acta》1988,148(2):233-240
The complexes CodptX3 and [Codpt(H2O)X2]ClO4 (X = Cl, Br; dpt = dipropylenetriamine = NH(CH2CH2CH2NH2)2) have been prepared and characterized. Rate constants (s−1) for aqueous solution at 25 °C and μ = 0.5 M (NaClO4), for the acid-independent sequential ractions.
have been measured spectrophotometrically. For X = Cl: k1 ⋍ 2 × 10−2, k2 = 1.7 × 10−4 and k3 = 4.8 × 10−6, and for X = Br: k1 ⋍ 2 × 10−2, k2 = 5.25 × 10−4 and k3 = 2.5 × 10−5 The primary equation was found to be acid independent, while the secondary and tertiary aquations were acid-inhibited reactions. For the second step, the rate of the reaction was given by the rate equation
where Ct is the complex concentration in the aqua-and hydroxodihalo species, k2 is the rate constant for the acid-dependent pathway and Ka is the equilibrium constant between the hydroxo and aqua complex ions. The activation parameters were evaluated, for X = Cl: ΔH2 = 106.3 ± 0.4 kJ mol−1 and ΔS2 = 40.2 ± 1.7 J K−1 mol, and for X = Br: ΔH2 = 91.6 ± 0.4 kJ mol−1 and ΔS2 = 0.4 ± 1.7 J K−1 mol−1. The results are discussed and detailed comparisons of the reactivities of these complexes with other haloaminecobalt(III) species are presented.  相似文献   

4.
The preparation of the planar yellow [Ni([8]aneN2)2](ClO4)2 is described. The complex dissociates in basic solution, with rate = kOH[NiL][OH?] (L = 1,5-diazacyclo-octane). At 25 °C, kOH = 4.5 x 10?2 M?1 s?1 and the corresponding activation parameters are ΔH = 69.2 kJ mol?1 and ΔS298 = ?38.6 J K?1 mol?1. Acid catalysed dissociation in quite slow even in strongly acidic solutions. The kinetic data in this case can be fitted to the expression Kobs = ko + KH[H+], where ko relates to a solvolytic pathway and kH to the acid catalysed pathway. At 60 °C, Ko = 2 x 10?5 s?1 and kH is 2 x 10?5 M?1 s?1. Possible mechanisms for these reactions are considered.The Ni(II)/Ni(III) redox couple for NiLn+ is irreversible on Pt using MeCN as solvent.  相似文献   

5.
Thermodynamic parameters for the reduction of ferrioxamine E as calculated from redox potentials determined at four different temperatures were found to be ΔH=7.1±3.4 kJ mol?1 and ΔS=?146 J mol?1 K?1. The negative entropy value is large, because the decrease in the charge at the metal center and an increase in its ionic radius force the structure of the complex to become less rigid and resemble the desferrisiderophore. The hydrophilic groups of the system are now (relatively more) available for solvent interaction. Thus, a large negative entropy change accompanies the reduction of the complex. Kinetics of reduction of ferrioxamine by VII, CrII, EuII, and dithionite were measured at different temperatures and by dithionite at different pH values. The CrII and EuII reactions proceed by an inner‐sphere mechanism and have second‐order rate constants at 25° of 1.37×104 and 1.23×105 M ?1 s?1, respectively. For the VII reduction, the corresponding rate constant was 1.89×103 M ?1 s?1. The activation parameters for the VII reduction were ΔH = 8.3 kJ mol?1; ΔS = ?154 J mol?1 K?1. These values are indicative of an outer‐sphere mechanism for VII reduction. The reduction by dithionite is half order in dithionite concentration indicating that SO . is the sole reducing species. log of reduction rate constants of different trihydroxamates by this reductant were correlated with their respective redox potentials, and the variation was found to be in approximate correspondence with the expectations of Marcus relationship.  相似文献   

6.
Using a liquid chromatography method that separates the two sulfonium diastereoisomers of adenosylmethionine, we have found that immature soybeans, soybean callus culture, radish leaves, yeast and rat liver contain only the (S)-sulfonium form of S-adenosylmethionine. Our findings contradict the suggestion by Stolowitz and Minch that 10–20% of naturally-occurring adenosylmethionine may have the (R)-configuration at the sulfonium pole. Absence of the (R)-sulfonium isomer of adenosylmethionine in biological materials indicates that the (R)-sulfonium form of adenosylmethionine present in commercial adenosylmethionine samples is an artifact of the isolation procedure. Our method of measuring the isomers of adenosylmethionine enabled us to readily determine the rate of racemization and hydrolysis of adenosylmethionine. Our rate constants for racemization (Kr) and hydrolysis (Kh) were 2.4 × 10?6 sec?1 and 12.3 × 10-?6 sec?1, respectively; values which are noticeably different from those of Wu and co-workers which were obtained with a more complicated method (Kr = 8 × 10?1 sec?1; Kh = 6 × 10?6 sec?1). We believe the absence of the (R)-isomer in vivo is best explained by stabilization of the (S)-isomer as suggested by Wu et al. Although the tissues we have analysed contained the (S)-sulfonium form of adenosylmethionine exclusively, when ethionine-resistant soybean cell lines were given ethionine, they accumulated both sulfonium diastereoisomers of adenosylethionine.  相似文献   

7.
The binding of Cu2+ to apostellacyanin occurs in two steps. The first step consists of a fast equilibrium reaction involving binding of copper to the protein in a non-native, though specific way, as shown by electron paramagnetic resonance measurements. All the spectroscopic properties of native stellacyanin are recovered in a slower monomolecular process (k = 7.5 × 10?3 sec?1 at 25 °C) characterized by high activation energy (ΔHa = 22 kcal mole?1) and low activation entropy (ΔSa = 3.0 cal deg?1 mole?1). The second step parallels a conformational change of the copper-bound protein molecule. A large difference of the tyrosyl residues pKs is found between holo- and apostellacyanin. In the latter the tyrosyl residues appear to be more exposed to solvent perturbations. Ammonia or monovalent anions such as N3?, SCN?, and Cl? have a catalytic effect on the second step of the reaction, roughly proportional to their first binding constant to aqueous copper. It is suggested that they may compete for a non-native bond of the copper to the protein, thus rendering the conformational change easier.The effect of Ag3 and Hg2+ on the recombination reaction with copper is discussed in terms of conformation of the metal-bound protein.  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.
Rate parameters have been obtained for the oxidation of cuprous stellacyanin by cobalt(III) ions of the form cis(N)-[CoN2O4]?, including cis(N)-[Co(NTA)(gly)]?, cis(N)-[Co(IDA)2]?, [Co(en)(ox)2]?(μ 0.5 M(phosphate), pH 7.0), and Co(EDTA)?(μ 0.1 M(NaCl), pH 7.2, 0.001 M phosphate). An excellent isokinetic correlation between the activation parameters ΔH and ΔS exists for the reactions of aminopolycarboxylatocobalt(III) ions with reduced stellacyanin (β = 300 ± 12 K; correlation coefficient = 0.995). It is concluded that enthalpy-entropy compensation in these reactions may be understood in terms of differing orientations preferred by the various oxidants in forming precursor complexes with the reduced blue protein. While ΔH and ΔS values for electron transfer from stellacyanin to cis(N)-[CoN2O4]? ions vary over ranges of 10.7 kcal/mol and 34 cal/mol-deg, respectively, room temperature rate constants are relatively constant (3.6–34.5 M?1 sec?1), as expected from Marcus theory for outer sphere electron transfer.  相似文献   

10.
Calcium-binding stoichiometry, dissociation equilibrium constants at zero ionic strength (K0), and molar extinction difference coefficients (Δ?λ) at the wavelength λ of the metallochromic indicators arsenazo I (ArsI) and tetramethylmurexide (TMX) were reevaluated with a computerized method based on mass conservation and thermodynamic consistency checks. This new method is shown to provide a more critical assessment of the assumed calcium-dye complexing model than is afforded by the commonly used reciprocal-plot method. The analyses of spectrophotometric Ca titrations confirm that both dyes form only 1:1 complexes in aqueous solution. For TMX, K0 = 1.3 × 10?3m and Δ?480 = 1.5 × 104m?1 cm?1; for ArsI, K0 = 5.8 × 10?3m and Δ?562 = 1.8 × 104m?1 cm?1 at pH 7.0 and T = 293°K. The discriminatory power of the analytical method is demonstrated by comparison of these results with those found for a different dye, arsenazo III, which complexes Ca in 1:1, 1:2, and 2:1 forms.  相似文献   

11.
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   

12.
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   

13.
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   

14.
Detailed analyses of changes in the ultraviolet-visible absorption spectra of the anti-aithritic gold drug disodium gold(I) thiomalate·0 3 glycerol·2H2O with time, suggest that the solid may contain about 23% of a species with λmax of 337 and 370 nm. This disappears in a two-step process soon after dissolution in water. The reaction was monitored at a variety of temperatures (20–47°C), pH's (6–11), and ionic strengths (0.05–0.61 M). The first step is complete in ca. 3 min. The second step is independent of Au(tm) concentration with ko' = 8.5 × 10?2min?1 and activation parameters of ΔH± = 82.1 4.1 kJmol?1 and ΔS = 13.65 KJ?1 mol?1. The logarithm of the rate of this step increases linearly with the square root of the ionic strength. The reaction is readily reversed at high ionic strengths and is interpreted as a cooperative structural transition of polymeric gold(I) thiomalate, possibly involving Au(I)-Au(I) bonding. The relationship of these observations to reactions of other 1:1 Au(I) thiolate complexes and their method of preparation is discussed.  相似文献   

15.
The kinetics of malonate replacement in bis- (malonato)oxovanadate(IV), [VO(mal)2H2O]2−(hereafter water molecule will be omitted), by oxalate has been studied by the stopped-flow method. The reaction was found to consist of two consecutive steps (k1 and k2: first-order rate constants) passing through a mixed ligand complex, [VO(mal)(ox)]2−. The rates for each step depended linearly on the concentrations of free oxalate species, Hox and ox2−. The second-order rate constants for the replacement by ox2− were much larger in the k1 step than in the k2 step and the activation parameters were determined as follows: ΔH= 43.5 ± 5.6 kJ mol−1, ΔS±-53 ± 19 J K−1 mol−1 and ΔH≠= 43.6 ± 0.5 kJ mol−1, δS≠ = -62 ± 2 J K−l mol−1 for the k1 and k2 steps, respectively. The volume of activation was determined to be -0.65 ± 0.75 cm3 mol−1 at 20.2 °C by the high-pressure stopped-flow method for the apparent rate constants.  相似文献   

16.
The binding of cis(c)- and trans(t)-Pt(NH3)2Cl2 to DNA at platinum/DNA-nucleotide ratios (Ri) of 0.1 or less has been studied by means of radioactive 195mPt-labeled compounds. Kinetic data are consistent with the following scheme:
At 25°C and pH 5–6 in 5 mM NaClO4, the values for the rate constants in the above scheme for the c-isomer are k2 = 2.2 × 10?5 sec?1, k7 = 0.32 (sec M)?1, and k8 = 143 (sec M)?1; for the t-isomer the values are k2 < 0.5 × 10?5 sec?1 and k7 = 0.95 (sec M)?1. Platinum-DNA adducts do not undergo detectable exchange after 3 days at 37°C, indicating the absence of a dynamic equillibrium. For both isomers the rate of binding is the same for single- and double-stranded DNA. The conclusions derived from Ag+ and H+ titration studies are consistent with binding at guanine N(7) for Ri < 0.1. The reaction rate is competitively inhibited by various salts and buffers and is suppressed by raising the pH (50% inhibition of initial rates at pH 7.3). At 37°C and pH 7 in 0.15 M NaCl, 6–8% of both the c- and t-isomers bind to DNA in 24 h, suggesting that both compounds should bind to DNA under biological conditions.  相似文献   

17.
《Inorganica chimica acta》1988,149(2):259-264
The bis(N-alkylsalicylaldiminato)nickel(II) complexes Ni(R-sal)2 with R = CH(CH2OH)CH(OH)Ph (I), R = CH(CH3)CH(OH)Ph (II) and R = CH2CH2Ph (III; Ph = phenyl) were prepared and characterized. In the solid state I and II are paramagnetic (μ = 3.2 and 3.3 BM at 20 °C, respectively), whereas III is diamagnetic. It follows from the UV-Vis spectra that in acetone solution I is six-coordinate octahedral and III is four-coordinate planar, the spectrum of II showing characteristics of both modes of coordination. Vis spectrophotometry and stopped-flow spectrophotometry were applied to study the kinetics of ligand substitution in I–III by H2salen (= N,N′-disalicylidene-ethylenediamine) in the solvent acetone at different temperatures. The kinetics follow a second-order rate law, rate = k[H2-salen] [complex]. At 20 °C the sequence of rate constants is k(III):k(II):k(I) = 11 850:40.6:1. The activation parameters are ΔH(I) = 112, ΔH(II) = 40.7, ΔH(III) = 35.7 kJ mol−1 and ΔS(I) = 92, ΔS(II) = −103, ΔS(III) = −89 J K−1 mol−1. The enormous difference in rate between complexes I, II and III, which is less pronounced in methanol, is attributed to the existence of a fast equilibrium planar ⇌ octahedral, which is established in the case of I and II by intramolecular octahedral coordination through the hydroxyl groups present in the organic group R. An A-mechanism is suggested to control the substitution in the sense that the entering ligand attacks the four-coordinate planar complex, the octahedral complex being kinetically inert.  相似文献   

18.
19.
The serine protease enteropeptidase exhibits a high level of substrate specificity for the cleavage sequence DDDDK~ X, making this enzyme a useful tool for the separation of recombinant protein fusion domains. In an effort to improve the utility of enteropeptidase for processing fusion proteins and to better understand its structure and function, two substitution variants of human enteropeptidase, designated R96Q and Y174R, were created and produced as active (>92%) enzymes secreted by Pichia pastoris with yields in excess of 1.7 mg/Liter. The Y174R variant showed improved specificities for substrates containing the sequences DDDDK (kcat/KM = 6.83 × 106 M?1 sec?1) and DDDDR (kcat/KM = 1.89 × 107 M?1 sec?1) relative to all other enteropeptidase variants reported to date. BPTI inhibition of Y174R was significantly decreased. Kinetic data demonstrate the important contribution of the positively charged residue 96 to extended substrate specificity in human enteropeptidase. Modeling shows the importance of the charge–charge interactions in the extended substrate binding pocket.  相似文献   

20.
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   

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