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.
Calorimetric studies of the reduction of free oxygen in solution by sodium dithionite are in agreement with a stoichiometry of 2 moles Na2S2O4 per mole of oxygen. The reaction is biphasic with ΔHt - 118±7 kcal mol?1 (?494 ± 29 kJ mol?1). The initial phase of the reaction proceeds with an enthalpy change of ca ?20 kcal (?84 kJ) and occurs when 0.5 moles of dithionite have been added per mole dioxygen present. This could be interpreted as the enthalpy change for the addition of a single electron to form the superoxide anion. Further reduction of the oxygen to water by one or more additional steps is accompanied by an enthalpy change of ca ?100 kcal (?418. 5 kJ). Neither of these reductive phases is consistent with the formation of hydrogen peroxide as an intermediate. The reduction of hydrogen peroxide by dithionite in 0.1 M phosphate buffer, pH 7.15, is a much slower process and with an enthalpy change of ca ? 74 kcal mol?1 (?314 kJ mol?1). Dissociation of oxyhemoglobin induced by the reduction of free oxygen tension with dithionite also shows a stoichiometry of 2 moles dithionite per mole oxygen present and an enthalpy change of ca. ?101 ±9 kcal mol?1 (?423± 38 kJ mol?1). The difference in the observed enthalpies (reduction of dioxygen vs. oxyhemoglobin) has been attributed to the dissociation of oxyhemoglobin, which is 17 kcal mol?1 (71 kJ mol?1).  相似文献   

2.
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   

3.
Perutz & Brunori (1982) proposed that the COOH-terminal His and Ser F9 of the beta-chains of fish and amphibian hemoglobins are responsible for their Root effect and part of their alkaline Bohr effect. Analysis of the kinetics of carbon monoxide binding by hemoglobin from the tadpole of Xenopus laevis supports that model and suggests an explanation for the absence of an alkaline Bohr effect in many aquatic Anura and Urodela.  相似文献   

4.
Kinetic evaluation of the oxidation of oxymyoglobin (MbO2) to metmyoglobin (Mb+) by bis(dimethylglyoximato)cobalt nitrosyl [Co(NO)(DMGH)2] has established that the mechanism of this transformation involves initial dissociation of nitric oxide from Co(NO)(DMGH)2, followed by direct oxidation of MbO2 by nitric oxide. Nitrate formation accompanies the production of Mb+ and is proposed to arise from isomerization of the initially formed peroxynitrite ion. By comparative kinetic determinations with nitrosyl transfer from the cobalt nitrosyl reagent to deoxyhemoglobin, the rate constant for oxidation of MbO2 by nitric oxide is calculated to be 31 X 106 M?1sec?1 at 10.0°C in phosphate-buffered media at pH 7.0. Bis(dimethylglyoximato)cobalt(II), the cobalt complex formed by nitric oxide dissociation from Co(NO)(DMGH)2, is an effective trap for dioxygen liberated from MbO2. The resulting μ-peroxo- or μ-superoxo-dicobaloxime(III) oxidizes deoxymyoglobin to metmyoglobin at a rate that is competitive with oxidation induced by Co(NO)(DMGH)2.  相似文献   

5.
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   

6.
Yeast enolase binds one mole of 113Cd2+ per subunit at a site that consists of all oxyligands in a distorted octahedral environment. This "conformational" metal ion's environment undergoes further distortion on addition of substrate/product or analogs. At pH's below the optimum value the shifted resonance tends to break up into several, suggesting the existence of several slowly exchanging intermediate forms. At acid pH's, on addition of one additional mole/subunit of 113Cd2+, which greatly increases catalysis, "conformational" resonance(s) further broadens, suggesting that the second, "catalytic" metal ion increases the rates of interconversion between "conformational" species. At more alkaline pH's, near the optimum pH, the "conformational" peak is sharpened, which suggests that very fast interconversion is occurring. The position of the "catalytic" metal ion resonance also suggests all oxyligands in a distorted octahedral geometry. The "catalytic" resonance is often broadened to the point where it cannot be seen, suggesting rapid changes in its geometry due to interconversion of substrate and product.  相似文献   

7.
Activation of yeast enolase by Cd2+ exhibits properties similar to activation by the physiological cofactor Mg2+. The activity is weakly stimulated, then inhibited by increasing ionic strength. The activity increases, then falls with increasing Cd2+ concentration. The effect of pH on activity produced by Cd2+ is very similar to that produced by Mg2+, except that the Cd2+ profile is shifted one pH unit to more alkaline values, and the maximum activity of the Cd2+-enzyme is about 10% of that of the Mg2+-enzyme. The apparent kinetic parameters of Cd2+ activation show little effect of pH except for inhibition by high concentrations of Cd2+: the apparent Ki increases sharply with pH. This is interpreted as the result of Cd2+ being a less effective "catalytic" metal ion, and Cd2+ being more effective in stabilizing the enzyme at alkaline pH's. The similarity of effects of ionic strength, divalent cation, and pH may be due to interaction with the same six sites per mole of enzyme. We also characterized the dependence of what is believed to be the enzyme-catalyzed enolization of a substrate analog, D-tartronate semialdehyde-2-phosphate (TSP) on similar parameters of pH, ionic strength, etc. The putative enolization is dependent on catalytic metal ion, although the TSP binds to the conformational Cd2+-enzyme complex. The reaction is very slow and very pH dependent, increasing with pH with a midpoint of reaction velocity at pH 8.7. There is a strong qualitative correlation between pH dependencies of reaction velocity of substrate conversion and TSP enolization and absorbance of the enzyme-bound TSP enolate, whether with Mg2+ or Cd2+ as cofactor. The slowness of the Cd2+-TSP reaction is not limited by proton release or any reaction involving covalent bonds to hydrogen. The apparent reaction rate constant increases linearly with Cd2+ concentration. Addition of excess ethylenediaminetetraacetic acid reverses the TSP reaction, but again very slowly. The binding of Cd2+ to the catalytic sites is characterized by low association and dissociation rate constants.  相似文献   

8.
The iron(III) complex of deuteroporphyrin(IX), deuteroferriheme, catalyzes the chlorination, by sodium chlorite, of the active methylene compound monochlorodimedone (MCD) to dichlorodimedone. Rate studies, carried out on a stopped-flow spectrophotometric time scale, show the chlorination to be zero-order in MCD, first-order in ClO2- and to display a complex dependence on heme. The active chlorinating agent is believed to be hypochlorite, OCl-, formed as a result of the initial two-electron oxidation of heme to peroxidatic intermediate by chlorite ion. This scheme is supported by the fact that the normal (4:1) heme:ClO2- molar stoichiometry is reduced in the presence of MCD to values approaching 2:1. This suggests that MCD is an effective scavenger of OCl-, which, in the absence of active methylene compound, serves as a two-electron oxidant of heme. The zero-order dependence of rate on MCD is attributed to the slow formation of OCl-, consequent to a mechanism in which the rate-limiting step is viewed to be the regeneration of free heme from peroxidatic intermediate, probably via a catalatic pathway. Support for such a mechanism is provided by the fact that addition of ascorbate greatly enhances the rate of MCD chlorination, presumably by accelerating the rate of heme regeneration via perioxidation reduction of the heme intermediate.  相似文献   

9.
Two novel complexes of Pd(II) involving vitamin B6 compounds have been synthesized. They are compatible with the compositions Pd(P.H.)2 C2(P=pyridoxol) and Pd(PL.H)2C2 (PL = pyridoxal). The complexes inhibited the growth as well as the biosynthesis of RNA, DNA, and protein of E. coli B-766. Photoacoustic spectral (PAS) measurements showed that the complexes bound to DNA of the bacteria and were present only in the kidney of treated mice. The complexes inhibited the incorporation of 3H-thymidine as well as 14C-leucine in the DNA and protein, respectively, of liver cell cultures (BL8L). The inhibition of cell division of Walker-S-cells and human lymphocytes by the complexes was highly significant.  相似文献   

10.
Ultraviolet difference spectroscopy has been used to study Fe (III)-apoferritin complexes formed after addition of Fe (II) to apoferritin in air. At constant iron, the recorded spectra varied with time after Fe (II) addition and with the number of iron atoms/molecule (protein concentration). The results indicate that after production of an initial complex, rearrangement or migration of Fe (III) atoms occurs, with polynuclear species forming as end-product, probably by hydrolytic polymerization. The presence of Tb3+ or Zn2+ ions affected the Fe (III) spectra and their development in different ways. The combined data suggest that more than one site, or processes, are involved in ferritin iron-core formation and that some of the metal sites are clustered.  相似文献   

11.
Co(ethylenethiourea)2(CH3CO2)2, with a pseudo-tetrahedral structure but having a long-bonded fifth ligand, gives an esr spectrum almost identical with that of Co-Bovine Carbonic Anhydrase-OH; a previous assignment of a similar structure to the CoII-containing enzyme is then substantiated. Several other tetrahedral and trigonal-bipyramidal CoII complexes have also been investigated. Although none gives δ values close to those of CoBCAOH and the correlation high δ = trigonal bipyramidal, low δ = tetrahedral seems to hold, the influence of even minor distortions within pseudo-tetrahedral symmetry on the esr spectra indicates that, used alone, esr spectra may not be an unambiguous structure marker for CoII enzymes.  相似文献   

12.
The preparation and molecular and crystal structure of the complex [(ethylenediamine)bis(7,9,-dimethylhypoxanthine)platinum(II)] hexafluorophosphate, [Pt(C2H8N2)(C7H8N4O)2] (PF6)2, are reported. The complex crystallizes in the monoclinic system, space group C2/c, with a = 12.334(2)Å, b = 10.256(2)Å, c = 22.339(3)Å, β = 101.31(1)°, V = 2771.0Å3, Z = 4, Dmeasd = 2.087(3) g cm?3, Dcalc = 2.094 g cm?3. Intensities for 3992 symmetry-averaged reflections were collected in the θ-2o scan mode on an automated diffractometer employing graphite-monochromatized MoKα radiation. The structure was solved by standard heavy-atom Patterson and Fourier methods. Full matrix least-squares refinement led to a final R value of 0.051. Both the ethylenediamine chelate and the PF6? anion are disordered. The primary coordination sphere about the Pt(II) center is approximately square planar with the bidentate ethylenediamine ligand and the N(1) atoms [Pt(II) ? N(1) = 2.020(5)Å] of two 7,9-dimethylhypoxanthine bases (related by a crystallographic twofold axis of symmetry) occupying the four coordination sites. The exocyclic O(6) carbonyl oxygen atoms of the two 7,9-dimethylhypoxanthine ligands participate in intracomplex hydrogen bonding with the amino groups of the ethylenediamine chelate [N(ethylenediamine) ? O(6) = 2.89( )Å]. The observed Pt ? O(6) intramolecular distances of 3.074(6)Å are similar to those found in other Pt(II) N(1)-bound 6-oxopurine complexes and in several Pt(II) N(3)-bound cytosine systems.  相似文献   

13.
The chelating behavior of a new multidentate ligand with tuberculostatic activity toward Cu(II) and Ni(II) ions has been studied. This ligand 3-(2-carboxyhydrazine)phenylimino-2-oximobutane(H2C POB) is found to chelate the above metal ions in both its keto and enol forms. The probable structures of all the complexes and the location of the bonding sites have been established through magnetic and spectroscopic (infrared, electronic) studies. The Cu(II) complex of the enol form exhibits subnormal magnetic moment at room temperature, indicating the probable existence of some sort of super exchange phenomenon in the system. The ligand itself and a few of its Cu(II) complexes have been found to exert powerful in vitro antibacterial activity toward some tuberculosis mycobacteria, such as Mycobacterium flae, Mycobacterium smegmatis, and Mycobacterium H37Rv.  相似文献   

14.
Various His-Pt(II) coordination compounds were prepared by reaction of K2PtCl4 or cis-[Pt(NH3)2Cl2](cis-DDP) with His and analyzed by 1H and 13C NMR spectroscopy, electrophoresis, and ion-exchange chromatography. His may be coordinated to Pt by the imidazol iminogroup and/or the α-aminogroup; the carboxy group remains always free. Both bidentate as well as monodentate ligands were identified. Cis-DDP reacts with His to give a mixture of compounds where all these possibilities are present: cis-diamine-(histidine-N,N-)Pt(II) and three different types of cis-diammine-bis(histidine). HCl trans cleavage of compounds with bidentate His ligands leads to a mixture of two compounds having His ligated to Pt by an amino or imin group. The methods applied are suitable for analyzing reactions of His with cis-DDP under model conditions similar to physiological conditions.  相似文献   

15.
Syntheses and the results of a potentiometric and spectrophotometric study at 25°C and an ionic strength of 0.10 mol dm?3 (KNO3) of the H + and Cu2+ complexes of two pairs of tetra- and octapeptides are reported. The peptides, Gly-Gly-Pro-Gly and (Gly-Gly-Pro-Gly)2, and Gly-Gly-Pro-Lys and (Gly-Gly-Pro-Lys)2, are models for the biologically active octapeptide with antitumor activity, dog tuftsinyltuftsin. The results clearly demonstrate participation of the ?-NH2 group of -Lys- in metal ion coordination from below pH 6 and, with the octapeptides at high pH, suggest the presence of a bent conformation with a large chelate ring spanning the two ends of the peptide chain.  相似文献   

16.
The mechanisms by which histidine stabilizes the cobalt(II)-carnosine complex from oxidation to cobalt(III) in aqueous solution are investigated with 1H-nmr, laser Raman, and Fourier transform-infrared spectroscopy. Histidine has at least three effects on the cobalt(II)-carnosine complex. First, over the concentration range of at least 5 to 250 mM, histidine stabilizes the cobalt(II)-carnosine complex from oxidation by excluding solvent molecules from the equatorial coordination positions. Second, at the upper end of this concentration range, histidine reduces the strained nonplanarity of the equatorial coordination positions around the cobalt(II) ion that results from tridentate chelation by carnosine. Bidentate ligation by histidine causes the carnosine to bind as a bidentate ligand also. Third, bidentate ligation of two carnosine molecules to the equatorial coordination positions of Co(II) ion places the β-alanyl residues inthe vicinity of the two axial coordination positions and thereby inhibits the binding of molecular oxygen. Substitution of a molecule of histidine for one of these two carnosine molecules makes an axial coordination position available for binding oxygen. The first two effects are expected to stabilize the cobalt(II) ion from rapid oxidation, whereas the third effect is expected to give long-term stability of the peroxo-bridged complex. Since bidentate ligation of histidine is favored over monodentate ligation only when the concentration of Co(II) ion is not limiting and is inhibited by high concentrations of carnosine in the same solution, the results presented provide a possible explanation for the observation that the stability of the Co(II) complexes toward oxidation and their ability to bind molecular oxygen depend on both the relative and absolute concentrations of Co(II) ion, carnosine, and histidine in solution. Furthermore, these results provide additional support to the suggestion that the high activity of carnosinase in kidney is involved in part in regulation of the oxygen sensor in this organ.  相似文献   

17.
Complexation in the H+-Si(OH)4-tropolone (HL) system was studied in 0.6 M (Na)Cl medium at 25° C. Speciation and formation constants were determined from potentiometric (glass electrode) and 29Si-NMR data. Experimental data cover the ranges 1.5 ? - log[H+] ? 8.4, 0.002 ? B ? 0.012 M, and 0 ? C ? 0.060 M (B and C stand for the total concentration of Si and tropolone, respectively). In acid solutions (-log[H+] ? 3) a hexacoordinated cationic complex, SiL3+, is formed with log K(Si(OH)4 + 3HL + H+ XXX SiL3+ + 4H2O) = 7.08 ± 0.03. Furthermore, the formation of a disilicic acid was established from 29Si-NMR data. The dimerization constant of Si(OH)4 was found to be 10 exp (1.2 ± 0.1). In model calculations the solubility of quartz and amorphous SiO2 in the presence of tropolone is demonstrated. Data were analyzed using the least-squares computer program LETAGROPVRID.  相似文献   

18.
The reaction of 8-thioguanosine (8-thioGuoH2 with methylmercury(II) has been shown to give rise to 1:1 (neutral and cationic), 1:2 (neutral and cationic), and 1:3 (cationic) complexes of the type [CH3Hg(8-thioGuoH)], [(CH3Hg(8-thioGuoH2)]NO3, [(CH3Hg)2(8-thioGuo)], [(CH3Hg)2(8-thioGuoH)]NO3 and [(CH3Hg)3(8-thioGuo)]NO3, depending upon the reactant stoichiometry and pH. 1H NMR, 13C NMR, and IR, as well as analytical data were used to characterize the complexes. Coupling of methylmercury(II)-protons to mercury-199 has been observed in all compounds. The magnitude of the coupling, 2J(1H-199Hg), is strongly dependent on the nature of the ligand bonded to the methylmercury(II) group and correlates with the 13C chemical shifts of coordinated CH3Hg(II) at the different binding sites.  相似文献   

19.
Saturation kinetics are observed in the inhibition of cobalt carboxypeptidase A by the chelating agent 1,10-phenanthroline. The association constant K1 for the formation of the enzyme-metal-ligand ternary complex and k2, the rate of breakup of the ternary complex, have been obtained. A mechanism is proposed to account for the pH profile of the reaction which, in conjunction with K1, permits the calculation of the individual rate constants k1, K?1, k2, k3. The magnitude of the rate constant k1 suggests that cobalt(II) in CoCPA is five-coordinate. Similar but less extensive studies on inhibition by 2,2′-bipyridyl and 8-hydroquinoline-5-sulfonic acid have also been carried out  相似文献   

20.
Methylmercury(II) and mercury(II) complexes of imidazole (1), 1-methylimidazole (2), and the 1,3-dimethylimidazolium ion (3) have been prepared in aqueous or ethanolic solution. Elemental analysis and 1H nmr spectroscopy have been used to characterize the complexes. The MeHg (Me = methyl) binding sites have been identified as N1, N3 (1), N3, C2 (2), and C2 (3). Reaction with HgO leads to the formation of Hg-bridged complexes of the type Im-Hg-Im, (Im = imidazole), where bonding occurs through N1 (1) and C2 (3); the latter is also formed as a result of symmetrization of the C2-bound MeHg complex. The formation of the C2-bound (carbene) complexes is discussed in terms of the increased acidity of the C2 proton resulting from coordination of an electrophilic species at N3. Based on electrostatic considerations, there appears to be a “minimum degree of activation” required before C2 bonding can occur, which explains the lack of this coordination mode in 1. 199Hg-1H spin-spin coupling (4J) is observed for C-bound mercury, but not for N-bound mercury, which is interpreted in terms of a decreased ligand exchange rate in the former case, due to the greater stability of the Hg-C bond. 2J coupling constants measured in (CD3)2SO for a number of MeHg complexes of heterocyclic ligands (including the imidazoles of the present study) correlate well with the ligand pKa (25°C, aqueous solution), according to 2J = ?3.88 pKa + 248.5. Results in the present work are discussed in relation to our previous work with nucleosides. The significance of the results to biological systems is considered.  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