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 31 毫秒
1.
The association of imidazole and auramine O to native horse-liver alcohol dehydrogenase [Zn(II)LADH] and active-site specifically cobalt(II)-substituted horse-liver alcohol dehydrogenase [Co(II)LADH], respectively, has been investigated. In all cases [except imidazole binding to Zn(II)LADH in the presence of auramine O] the association rates approached an upper limit (kmax). The kmax values were compared for the metal ligands imidazole (monodentate), 1,10-phenanthroline and 2,2'-bipyridine (bidentate; see also the preceding paper), and for auramine O which does not coordinate to the catalytic metal ion. Independent of the large differences in their structure and metal-bonding capability, all these compounds exhibit common, maximum, limiting rate constants of about 60 s-1 and 200 s-1 for Co(II)LADH and Zn(II)LADH, respectively. These results demonstrate that kmax is strongly dependent on the catalytic metal ion but not on the ligand. The absence of spectral changes in the d-d transitions of the catalytic Co(II) ion upon auramine O binding to Co(II)LADH indicates that the rate-limiting step is not accompanied by a major conformational change. Finally, it is concluded that reactions in the inner coordination sphere of the catalytic metal ion (i.e. the metal-bound water molecule) are not responsible for the step characterized by kmax. We propose the rate-limiting step to consist of the dissociation of one or several water molecules from the second coordination sphere of the catalytic metal ion in the active site of LADH in its open conformation.  相似文献   

2.
1. The reactivity of the zinc site of bovine superoxide dismutase has been probed by observing optical and electron paramagnetic resonance changes, under several conditions, of the Co(II)-substituted protein. 2. Only in the absence of copper are the optical and electron paramagnetic resonance spectra of the cobalt chromophore appreciably affected by alkaline pH or by cyanide. With both reagents the reaction with the copper-containing protein appears to involve the water molecule bound to the copper and does not affect the magnetic coupling between copper and cobalt. 3. The reaction of cyanide with the copper-free Co(II) protein leads to a slow detachment of cobalt from the protein as pentacyanocobalt. An oxygen adduct forms in air, analogous to that described in Co(II) carbonic anhydrase (Haffner, P. H. and Coleman, J. E. (1975) J. Biol. Chem. 250, 996--1005.) 4. Acid titration modifies the Co(II) spectra in the same way in the Cu-containing and in the Cu-free protein and brings about uncoupling of the Co(II)--Cu(II) system. Protonation of histidine-61 on the zinc facing nitrogen is suggested. 5. H2O2 modifies the cobalt chromophore only in the presence of copper. Magnetic coupling between Cu(II) and Co(II) seems to be still present after H2O2 inactivation of the enzyme.  相似文献   

3.
D'souza VM  Bennett B  Copik AJ  Holz RC 《Biochemistry》2000,39(13):3817-3826
The metal-binding properties of the methionyl aminopeptidase from Escherichia coli (MetAP) were investigated. Measurements of catalytic activity as a function of added Co(II) and Fe(II) revealed that maximal enzymatic activity is observed after the addition of only 1 equiv of divalent metal ion. Based on these studies, metal binding constants for the first metal binding event were found to be 0.3 +/- 0.2 microM and 0.2 +/- 0.2 microM for Co(II)- and Fe(II)-substituted MetAP, respectively. Binding of excess metal ions (>50 equiv) resulted in the loss of approximately 50% of the catalytic activity. Electronic absorption spectral titration of a 1 mM sample of MetAP with Co(II) provided a binding constant of 2.5 +/- 0.5 mM for the second metal binding site. Furthermore, the electronic absorption spectra of Co(II)-loaded MetAP indicated that both metal ions reside in a pentacoordinate geometry. Consistent with the absorption data, electron paramagnetic resonance (EPR) spectra of [CoCo(MetAP)] also indicated that the Co(II) geometries are not highly constrained, suggesting that each Co(II) ion in MetAP resides in a pentacoordinate geometry. EPR studies on [CoCo(MetAP)] also revealed that at pH 7.5 there is no significant spin-coupling between the two Co(II) ions, though a small proportion ( approximately 5%) of the sample exhibited detectable spin-spin interactions at pH values > 9.6. EPR studies on [Fe(III)_(MetAP)] and [Fe(III)Fe(III)(MetAP)] also suggested no spin-coupling between the two metal ions. (1)H nuclear magnetic resonance (NMR) spectra of [Co(II)_(MetAP)] in both H(2)O and D(2)O buffer indicated that the first metal binding site contains the only active-site histidine residue, His171. Mechanistic implications of the observed binding properties of divalent metal ions to the MetAP from E. coli are discussed.  相似文献   

4.
W D Frasch  R Mei 《Biochemistry》1987,26(23):7321-7325
The evolution of O2 from H2O2 catalyzed by the oxygen-evolving complex (OEC) in darkness was examined with photosystem II reaction center complex preparations from spinach. Flash illumination of dark-adapted reaction centers was used to make S0-enriched or S1-enriched complexes. The membranes catalyzed O2 evolution from H2O2 when preset to either the S0 or S1 state. However, only the S0-state reaction was inhibited by carbonyl cyanide m-chlorophenylhydrazone and dependent on chloride. These results indicate that (1) the S0-dependent and S1-dependent catalytic cycles can be separated by flash illumination, (2) the S0-dependent reaction involves the formation of the S2 state, and (3) the S1-dependent reaction does not involve the formation of the S2 or S3 states. A kinetic study of the S1-dependent reaction revealed a rapid equilibrium ordered mechanism in which (1) the binding of Ca(II) must precede the binding of H2O2 to the OEC and (2) the reaction of Ca(II) with the free enzyme is at thermodynamic equilibrium such that Ca(II) does not necessarily dissociate after each catalytic cycle.  相似文献   

5.
In the site-specific mutant of human carbonic anhydrase in which the proton shuttle His64 is replaced with alanine, H64A HCA II, catalysis can be activated in a saturable manner by the proton donor 4-methylimidazole (4-MI). From 1H NMR relaxivities, we found 4-MI bound as a second-shell ligand of the tetrahedrally coordinated cobalt in Co(II)-substituted H64A HCA II, with 4-MI located about 4.5 A from the metal. Binding constants of 4-MI to H64A HCA II were estimated from: (1) NMR relaxation of the protons of 4-MI by Co(II)-H64A HCA II, (2) the visible absorption spectrum of Co(II)-H64A HCA II in the presence of 4-MI, (3) the inhibition by 4-MI of the catalytic hydration of CO2, and (4) from the catalyzed exchange of 18O between CO2 and water. These experiments along with previously reported crystallographic and catalytic data help identify a range of distances at which proton transfer is efficient in carbonic anhydrase II.  相似文献   

6.
J K Wright  J Feldman  M Takahashi 《Biochemistry》1976,15(17):3704-3710
The kinase active site of the aspartokinase-homoserine dehydrogenase enzyme complex of Excherichia coli has been affinity labeled both with substrates aspartate and adenosine triphosphate and feedback inhibitor threonine. Co(III) exchange-inert adducts of aspartokinase and inhibitor or substrates were produced in situ by oxidation of Co(II) with H2O2. Emzyme-Co(III)-adenosine 5'-triphosphate (ATP), enzyme-Co(III)-aspartate, and enzyme-Co(III)-threonine ternary adducts were produced in this manner. The formation of the enzyme-Co(III)-threonine adduct leads us to conclude that threonine inhibits the kinase activity of this enzyme complex by binding in the first coordination sphere of the catalytic metal ion cofactor, a conclusion which is consistent with evidence derived from previous nuclear magnetic resonance data obtained in this laboratory. The quaternary adducts formed by H2O2 oxidation in the presence of aspartokinase, Co(II), ATP, aspartate, and threonine comprised a mixture of both ezyme-Co(III)-ATP-aspartate and enzyme-Co(III)-ATP-threonine adducts. The formation of the quaternary aspartate-containing adduct was unexpected, since the presence of threonine was expected to prevent access of the aspartate to the active site; most significantly however, the the sum of the numbers of aspartate plus threonine molecules incorporated per active site is one. We believe that this shows direct steric overlap between the metal-adjacent binding sites for aspartate and threonine. Aspartate or threonine can not occupy the kinase active site simultaneously; this conclusion is consistent with the direct competitive inhibition of aspartate by threonine observed in steady-state kinetic studies.  相似文献   

7.
Design of chemically modified oligonucleotides for regulation of gene expression has attracted considerable attention over the past decades. One actively pursued approach involves antisense or antigene oligonucleotide constructs carrying reactive groups, many of these based on transition metal complexes. The complexes of Fe(II) and Co(II) with phthalocyanines are extremely good catalysts of oxidation of organic compounds with molecular oxygen and hydrogen peroxide. The binding of positively charged Fe(II) and Co(II) phthalocyanines with single- and double-stranded DNA was investigated. It was shown that these phthalocyanines interact with nucleic acids through an outside binding mode. The site-directed modification of single-stranded DNA by O2 and H2O2 in the presence of dimeric complexes of negatively and positively charged Fe(II) and Co(II) phthalocyanines was investigated. These complexes were formed directly on single-stranded DNA through interaction between negatively charged phthalocyanine in conjugate and positively charged phthalocyanine in solution. The resulting oppositely charged phthalocyanine complexes showed significant increase of catalytic activity compared with monomeric forms of phthalocyanines Fe(II) and Co(II). These complexes catalyzed the DNA oxidation with high efficacy and led to direct DNA strand cleavage. It was determined that oxidation of DNA by molecular oxygen catalyzed by complex of Fe(II)-phthalocyanines proceeds with higher rate than in the case of Co(II)-phthalocyanines but the latter led to a greater extent of target DNA modification.  相似文献   

8.
A series of heterobinuclear cofacial porphyrin-corrole dyads containing a Co(IV) corrole linked by one of four different spacers in a face-to-face arrangement with an Fe(III) or Mn(III) porphyrin have been examined as catalysts for the electroreduction of O(2) to H(2)O and/or H(2)O(2) when adsorbed on the surface of a graphite electrode in air-saturated aqueous solutions containing 1M HClO(4). The examined compounds are represented as (PCY)M(III)ClCo(IV)Cl where P is a porphyrin dianion, C is a corrole trianion and Y is a biphenylene (B), 9,9-dimethylxanthene (X), dibenzofuran (O) or anthracene (A) spacer. The catalytic behavior of the seven investigated dyads in the two heterobimetallic (PCY)MClCoCl series of catalysts is compared on one hand to what was previously reported for related dyads with a single Co(III) corrole macrocycle linked to a free-base porphyrin with the same set of linking bridges, (PCY)H(2)Co, and on the other hand to dicobalt porphyrin-corrole dyads of the form (PCY)Co(2) which were shown to efficiently electrocatalyze the four electron reduction of O(2) at a graphite electrode in acid media. Comparisons between the four series of porphyrin-corrole dyads, (PCY)Co(2), (PCY)H(2)Co, (PCY)FeClCoCl and (PCY)MnClCoCl, show that in all cases the biscobalt dyads catalyze O(2) electroreduction at potentials more positive by an average 110mV as compared to the related series of compounds containing a Co(III) or Co(IV) corrole macrocycle linked to a free-base metalloporphyrin or a metalloporphyrin with an Fe(III) or Mn(III) central metal ion. The data indicates that the E(1/2) values where electrocatalysis is initiated is related to the initial site of electron transfer, which is the Co(III)/Co(II) porphyrin reduction process in the case of (PCY)Co(2) and the Co(IV)/Co(III) corrole reduction in the case of (PCY)MnClCoCl, (PCY)FeClCoCl and (PCY)H(2)Co. The overall data also suggests that the catalytically active form of the biscobalt dyad in (PCY)Co(2) contains a Co(II) porphyrin and a Co(IV) corrole.  相似文献   

9.
The role of zinc in liver alcohol dehydrogenase has been studied by replacement of 1.3 and 3.5 of the four Zn(II) ions with Co(II) and measuring the effects of the paramagnetic Co(II) on the relaxation rates of the protons of water, ethanol, and isobutyramide. Water relaxation studies at 8, 24, 100, and 220 MHz indicate two classes of bound Co(II). The similar to 2 readily replaced Co(II) ions retain one fast exchanging water proton in their inner coordination spheres, while the similar to 2 slowly exchanging Co(II) ions coordinate no detectable water protons, indicating that the former replaced Zn(II) at the "catalytic sites" and the latter replaced Zn(II) at the "structural sites" detected crystallographically. Ethanol, acetaldehyde, and isobutyramide bind with appropriate affinities to the Co(II) substituted alcohol dehydrogenases decreasing the number of fast exchanging protons at the catalytic Co(II) site by greater than or equal to 54 percent. Coenzyme binding causes smaller changes in the water relaxation rate which may be due to local conformation changes. The paramagnetic effects of Co(II) at the catalytic site on the relaxation rates of the methyl protons of isobutyramide at 100 and 220 MHz indicate that this analog binds at a site 9.1 A from the catalytic Co(II). This distance decreases to 6.9 A when NADH is bound, and a Co(II) to methyne proton distance of 6.6 A is determined indicating a conformation change leading to the formation of a second sphere enzyme-Co(II)-isobutyramide complex in which a hydroxyl or water ligand intervenes between the metal and the substrate analog. Similar behavior is observed in the enzyme-ethanol complexes. The paramagnetic effects of Co(II), at the catalytic site, on the relaxation rates of the protons of ethanol at 100 and 220 MHz, indicate that this substrate bind at a site 12-14 A distant from the catalytic Co(II) but that this distancedecreases to 6.3 A in the abortive enzyme-NADH-ethanol complex. The role of the catalytic Co(II) thus appears to be the activation of a hydroxyl or water ligand which polarizes the aldehyde carbonyl group by hydrogen bonding. The role of the structural Co(II), which is more distant from isobutyramide (9-11 A), may be that of a template for protein conformation changes. By combining the present distances with those from previous magnetic resonance studies on the liver enzyme, the arrangement of coenzyme, metal, and substrate at the active site in solution can be constructed. This arrangement is consistent with that of ADP-ribose and zinc in the crystalline complex of liver alcohol dehydrogenase as determined by X-ray crystallography (Branden et al., (1973), Proc. Natl. Acad. Sci. U.S.A.70, 2439).  相似文献   

10.
The R2 dimer of mouse ribonucleotide reductase contains a dinuclear iron-oxygen cluster and tyrosyl radical/subunit. The dinuclear diferrous form reacts with dioxygen to generate the tyrosyl radical essential for the catalytic reaction that occurs at the R1 dimer. It is important to understand how the reactivity toward oxygen is related to the crystal structure of the dinuclear cluster. For the mouse R2 protein, no structure has been available with a fully occupied dinuclear metal ion site. A cobalt substitution of mouse R2 was performed to produce a good model for the very air-sensitive diferrous form of the enzyme. X-band EPR and light absorption studies (epsilon(550 nm) = 100 mm(-1) cm(-1)/Co(II)) revealed a strong cooperative binding of cobalt to the dinuclear site. In perpendicular mode EPR, the axial signal from mouse R2 incubated with Co(II) showed a typical S = 3/2 Co(II) signal, and its low intensity indicated that the majority of the Co(II) bound to R2 is magnetically coupled. In parallel mode EPR, a typical integer spin signal (M(s) = +/-3) with g approximately 12 is observed at 3.6 K and 10 K, showing that the two Co(II) ions (S = 3/2) in the dinuclear site are ferromagnetically coupled. We have solved the 2.4 A crystal structure of the Co(II)-substituted R2 with a fully occupied dinuclear cluster. The bridging Co(II) carboxylate ligand Glu-267 adopts an altered orientation compared with its counterpart Glu-238 in Escherichia coli R2. This might be important for proper O(2) activation of the more exposed native diferrous site in mouse R2 compared with E. coli R2.  相似文献   

11.
The aerobic oxidation of Co(II)bleomycin bound to calf thymus DNA has been investigated in relation to the mechanism of reaction in solution in the absence of DNA. Kinetics of dioxygenation of the Co(II) complex were followed by spectrophotometric and electron spin resonance spectroscopy as well as dioxygen analysis. The reaction is slower than when carried out in solution; its rate is inversely related to the ratio of DNA base pairs to Co(II)bleomycin. The subsequent oxidation reaction, observed spectrophotometrically and by dioxygen analysis, is second order in cobalt complex. The calculated second order rate constant is also inversely related to the base pair to metal complex ratio. Once this ratio exceeds three, the reaction rate slows significantly with each additional increment of DNA added to the starting reaction mixture. Taking advantage of the high stability of O(2)-Co(II)bleomycin bound to greater than a 3-fold excess of DNA base pairs, it could be demonstrated that the rate constant for oxidation of two O(2)-Co(II)bleomycin molecules is much slower than that for O(2)-Co(II)bleomycin plus Co(II)bleomycin. With the same technique it was observed that the metal centers of O(2)-Co(II)bleomycin and Fe(II)bleomycin also undergo oxidation. The binding to DNA of both solution products of the oxidation of Co(II)bleomycin by O2 was examined by 1H NMR spectroscopy. Peroxy-Co(III)bleomycin, Form I, binds with higher affinity than Co(III)bleomycin, Form II. At lower ionic strength, the size of the DNA binding site for each form is about 2 base pairs/molecule of drug.  相似文献   

12.
The aim of this study was to investigate the effect of dopa and catecholamines on the generation of oxygen active species during oxidation of farmorubicin by "Fenton-like" reagents using chemiluminescent and spectrophotometric techniques. The tested catechols were found to reduce the light emission accompanying oxidation of farmorubicin by Co(II) + H2O2 and Cu(II) + H2O2 mixtures. The quenching effect was followed by their rapid oxidation to aminochromes possessing toxic activities.  相似文献   

13.
Schwartz JK  Liu XS  Tosha T  Diebold A  Theil EC  Solomon EI 《Biochemistry》2010,49(49):10516-10525
DNA protection during starvation (Dps) proteins are miniferritins found in bacteria and archaea that provide protection from uncontrolled Fe(II)/O radical chemistry; thus the catalytic sites are targets for antibiotics against pathogens, such as anthrax. Ferritin protein cages synthesize ferric oxymineral from Fe(II) and O(2)/H(2)O(2), which accumulates in the large central cavity; for Dps, H(2)O(2) is the more common Fe(II) oxidant contrasting with eukaryotic maxiferritins that often prefer dioxygen. To better understand the differences in the catalytic sites of maxi- versus miniferritins, we used a combination of NIR circular dichroism (CD), magnetic circular dichroism (MCD), and variable-temperature, variable-field MCD (VTVH MCD) to study Fe(II) binding to the catalytic sites of the two Bacillus anthracis miniferritins: one in which two Fe(II) react with O(2) exclusively (Dps1) and a second in which both O(2) or H(2)O(2) can react with two Fe(II) (Dps2). Both result in the formation of iron oxybiomineral. The data show a single 5- or 6-coordinate Fe(II) in the absence of oxidant; Fe(II) binding to Dps2 is 30× more stable than Dps1; and the lower limit of K(D) for binding a second Fe(II), in the absence of oxidant, is 2-3 orders of magnitude weaker than for the binding of the single Fe(II). The data fit an equilibrium model where binding of oxidant facilitates formation of the catalytic site, in sharp contrast to eukaryotic M-ferritins where the binuclear Fe(II) centers are preformed before binding of O(2). The two different binding sequences illustrate the mechanistic range possible for catalytic sites of the family of ferritins.  相似文献   

14.
Manganese peroxidase (MnP), which normally oxidizes Mn2+ to Mn3+, is rapidly and completely inactivated in an H2O2-dependent reaction by 2 equivalents of sodium azide. The inactivation is paralleled by formation of the azidyl radical and high yield conversion of the prosthetic heme into a meso-azido adduct. The meso-azido enzyme is oxidized by H2O2 to a Compound II-like species with the Soret band red-shifted 2 nm relative to that of native Compound II. The time-dependent decrease in this Compound II-like spectrum (t1/2 = 2.3 h) indicates that the delta-meso azido heme is more rapidly degraded by H2O2 than the prosthetic heme of control enzyme (t1/2 = 4.8 h). MnP is also inactivated by phenyl-, methyl-, and ethylhydrazine. The phenylhydrazine reaction is too rapid for kinetic analysis, but KI = 402 microM and kinact = 0.22/min for the slower inactivation by methylhydrazine. Reaction with phenylhydrazine at pH 4.5 does not yield iron-phenyl, N-phenyl, or meso-phenyl heme adducts. Ethylhydrazine inactivates the enzyme both at pH 4.5 and 7.0, but only detectably produces delta-meso-ethyl-heme at pH 7.0. Reconstitution of apo-MnP with hemin or delta-meso-ethylheme yields enzyme with, respectively, 50 and 5% of the native activity. The delta-meso-alkyl group thus suppresses most of the catalytic activity of the enzyme even though a Compound II-like species is still formed with H2O2. Finally, Co2+ inhibits the enzyme competitively with respect to Mn2+ but does not inhibit its inactivation by azide or the alkylhydrazines. The results argue that substrates interact with the heme edge in the vicinity of the delta-meso-carbon. They also suggest that Mn2+ and Co2+ bind to a common site close to the delta-meso-carbon without blocking the approach of small molecules to the heme edge. An active site model is proposed that accommodates these results.  相似文献   

15.
C K Tu  D N Silverman 《Biochemistry》1985,24(21):5881-5887
We have measured the catalysis by Co(II)-substituted bovine carbonic anhydrase II from red cells of the exchange of 18O between CO2 and H2O using membrane-inlet mass spectrometry. We chose Co(II)-substituted carbonic anhydrase II because the apparent equilibrium dissociation constant of HCO3- and enzyme at pH 7.4, KHCO3-eff approximately equal to 55 mM, was within a practicable range of substrate concentrations for the 18O method. For the native, zinc-containing enzyme KHCO3-eff is close to 500 mM at this pH. The rate constant for the release from the active site of water bearing substrate oxygen kH2O was dependent on the fraction of enzyme that was free, not bound by substrate HCO3- or anions. The pH dependence of kH2O in the pH range 6.0-9.0 can be explained entirely by a rate-limiting, intramolecular proton transfer between cobalt-bound hydroxide and a nearby group, probably His-64. The rate constant for this proton transfer was found to be 7 X 10(5) S-1 for the Co(II)-substituted enzyme and 2 X 10(6) S-1 for the native enzyme. These results are applied to models derived from proton-relaxation enhancement of water exchanging from the inner coordination shell of the cobalt in carbonic anhydrase. The anions iodide, cyanate, and thiocyanate inhibited catalysis of 18O exchange by Co(II)-substituted carbonic anhydrase II in a manner competitive with total substrate (CO2 and HCO3-) at chemical equilibrium and pH 7.4. These results are discussed in terms of observed steady-state inhibition patterns and suggest that there is no significant contribution of a ternary complex between substrate, inhibitor, and enzyme.  相似文献   

16.
Two inhibitor-containing 'half-sandwich' cobalt(II) complexes [(TpPh)Co(X)(CH3OH)m] x nCH3OH ((TpPh) = hydrotris (3-phenylpyrazolyl)borate; 1: X- = N3-, m = 1, n = 2; 2: X- = NCS-, m = 0, n = 0) have been synthesized and used as the catalysts in the bicarbonate dehydration reaction. The structures of 1 and 2 were determined by X-ray diffraction analysis, which shows that N3- and NCS- coordinate to the Co(II) ions of 1 and 2, respectively, with the Co-N bond lengths of 1.992(6) A and 1.901(3) A. The coordination geometries of the Co(II) complexes in solution are five-coordinated trigonal bipyramid as revealed by the spectroscopic measurements. The dehydration kinetic measurements of HCO3- are performed by the stopped-flow techniques at pH < 7.9. The apparent dehydration rate constant k(obs) varies linearly with Co(II) complex and H+ concentrations, respectively, and the catalytic activity of 2 is lower than that of 1. The aqua Co(II) complex must be the reactive catalytic species in the catalyzed dehydration reaction and the rate-determining step is the substitution of the labile water molecule by HCO3-. The k(obs) values increase with increasing reaction temperature, and the large negative entropy of activation also indicates the associative activation mode. The inhibition ability of NCS- is stronger than that of N3-, which can be rationalized by the decreases in the Co-N(N3-/NCS-) bond lengths and effective atomic charges of the Co(II) ions based on the X-ray crystallographic data and theoretical calculations in this work.  相似文献   

17.
Enzyme elements that are involved in the reversible cyclization of L-carbamylaspartate to L-dihdroorotate catalyzed by dihydroorotase (EC 3.5.2.3) from Clostridium oroticum (ATCC 25750) have been studied. Removal of Zn(II) from the enzyme by chelators followed by incubation of apoenzyme with Co(II) results in replacement of two to three of the four Zn(II) ions per molecule by Co(II). The catalytic properties of the Zn(II)Co(II) dihydroorotase are different from those of native enzyme. The Vmax is increased for both the synthesis and hydrolysis of L-dihydroorotate. The Km for L-dihydroorotate is unchanged, while the Km for L-carbamylaspartate is increased more than twofold. On the other hand, the kinetic properties of Zn(II)-reconstituted dihydroorotase are indistinguishable from those of native enzyme. The pH dependence of Vmax is also altered by the Co(II) substitution. For both Zn(II)- and Zn(II)Co(II)-dihydroorotase, this pH dependence is well described by a single ionization and the pK's for L-dihydroorotate synthesis and hydrolysis are different. Substitution with Co(II) increases the pK for both reaction directions to different extents. These results strongly support a role for the tightly bound metals in the catalytic mechanism. In addition, diethylpyrocarbonate rapidly inactivates the enzyme. The inactivation is prevented by L-dihydroorotate. This result is consistent with a role for at least one histidine in catalysis. The possibility that C. oroticum dihydroorotase may be useful model for the more complex mammalian enzyme is considered.  相似文献   

18.
The synthesis, characterization and catalytic activity of a series of tetra-halogeno-dimethyl salen cobalt (II) complexes are reported in this paper. The investigated complexes of cobalt (II) with Schiff bases are: αα′-di-methyl Salen cobalt (II) [Co(dMeSalen)], 3,3′,5,5′-tetra chloro α,α′-di-methyl Salen cobalt (II), [Co(tCldMeSalen)], 3,3′-di-bromo 5,5′-di-chloro α,α′-di-methyl Salen cobalt (II), [Co(tBrdMeSalen)], 3,3′,5,5′-tetra bromo α,α′-di-methyl Salen cobalt (II), [Co(tBrdMeSalen)] and 3,3′,5,5′-tetra iodo α,α′-di-methyl Salen cobalt (II), [Co(tIdMeSalen)] (where Salen is bis(salicylaldehyde)ethylenediamine). The characterization of the complexes was performed by elemental analysis, cyclic voltammetry, UV-Vis, IR and EPR spectroscopies. The study was made in DMF, and pyridine was used for coordination as axial base. The redox potential is influenced by the substituent grafted on aromatic ring and in the azomethynic position and also by the molecules coordinating in axial position (solvent, DMF, or pyridine). The catalytic oxygenation of 2,6-di-tert-butylphenol by these complexes leads to the obtention of benzoquinone and diphenoquinone products. The cobalt (II) complexes form reversible adducts with molecular oxygen.  相似文献   

19.
Augmentation, by CO(2)/HCO(3)(-), of Co(II)-catalyzed peroxidations was explored to clarify whether the rate enhancement was due to CO(2) or to HCO(3)(-). The rate of oxidation of NADH by Co(II) plus H(2)O(2), in Tris or phosphate, was markedly enhanced by CO(2)/HCO(3)(-). Phosphate was seen to inhibit the Co(II)-catalyzed peroxidation, probably due to its sequestration of the Co(II). When CO(2) was used, there was an initial burst of NADH oxidation followed by a slower linear rate. The presence of carbonic anhydrase eliminated this initial burst; establishing that CO(2) rather than HCO(3)(-) was the species responsible for the observed rate enhancements. Both kinetic and spectral data indicated that Co(II) was converted by H(2)O(2) into a less active form from which Co(II) could be regenerated. This less active form absorbed in both the UV and visible regions, and is assumed to be a peroxy bridged binuclear complex. The rate of formation of this absorbing form was increased by HCO(3)(-)/CO(2). A minimal mechanism consistent with these observations is proposed.  相似文献   

20.
The title compounds, (cinchonidinium)trichlorocobalt(II) [(C(19)H(23)ON(2))CoCl(3)] (CoCdn) and (quinidinium)trichlorocobalt(II) [(C(20)H(25)O(2)N(2))CoCl(3)] (CoQd), are zwitterions that differ in absolute configuration and conformation. In both complexes, the sp(3) nitrogen of quinuclidine is protonated, whereas the sp(2) nitrogen of quinoline is linked to the Co(II) atom, which coordinates three chlorine atoms in distorted tetrahedral geometry. The mutual orientations of the quinoline and quinuclidine moieties in CoCdn and CoQd differ significantly because of different hydrogen bonding involving the hydroxyl group. In both complexes, the quinuclidine NH groups and hydroxyl groups are hydrogen-bond donors to the chlorine atoms of Co(II) tetrahedra. In CoQd the hydrogen bonding leads to formation of a nine-membered ring consisting of Co, two chlorines, and a fragment of the quinidine molecule. A comparison of the crystal structures of four Cinchona alkaloid complexes with trichlorocobalt(II) shows that their space groups are determined by the absolute configuration of the alkaloid, whereas the hydrogen-bonding pattern is mainly affected by the substituent in the quinoline ring, i.e., by hydrogen or methoxyl group.  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