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.
Flash photolysis of alkaline peroxynitrite solutions results in the formation of nitrogen monoxide and superoxide. From the rate of recombination it is concluded that the rate constant of the reaction of nitrogen monoxide with superoxide is (1.9 +/- 0.2) x 10(10) M-1 s-1. The pKa of hydrogen oxoperoxonitrate is dependent on the medium. With the stopped-flow technique a value of 6.5 is found at millimolar phosphate concentrations, while at 0.5 M phosphate the value is 7.5. The kinetics of decay do not follow first-order kinetics when the pH is larger than the pKa, combined with a total peroxynitrite and peroxynitrous acid concentration that exceeds 0.1 mM. An adduct between ONOO- and ONOOH is formed with a stability constant of (1.0 +/- 0.1) x 10(4) M. The kinetics of the decay of hydrogen oxoperoxonitrate are not very pressure-dependent: from stopped-flow experiments up to 152 MPa, an activation volume of 1.7 +/- 1.0 cm3 mol-1 was calculated. This small value is not compatible with homolysis of the O-O bond to yield free nitrogen dioxide and the hydroxyl radical. Pulse radiolysis of alkaline peroxynitrite solutions indicates that the hydroxyl radical reacts with ONOO- to form [(HO)ONOO].- with a rate constant of 5.8 x 10(9) M-1 s-1. This radical absorbs with a maximum at 420 nm (epsilon = 1.8 x 10(3) M-1 cm-1) and decays by second-order kinetics, k = 3.4 x 10(6) M-1 s-1. Improvements to the biomimetic synthesis of peroxynitrite with solid potassium superoxide and gaseous nitrogen monoxide result in higher peroxynitrite to nitrite yields than in most other syntheses.  相似文献   

2.
Earlier investigations of the oxidation of 3,5,3',5'-tetramethylbenzidine (TMB) using horseradish peroxidase and prostaglandin H-synthase have shown the formation of a cation free radical of TMB in equilibrium with a charge-transfer complex, consistent with either a two- or a one-electron initial oxidation. In this work, we exploited the distinct spectroscopic properties of myeloperoxidase and its oxidized intermediates, compounds I and II, to establish two successive one-electron oxidations of TMB. By employing stopped-flow techniques under transient-state and steady-state conditions, we also determined the rate constants for the elementary steps of the myeloperoxidase-catalyzed oxidation of TMB at pH 5.4 and 20 degrees C. The second-order rate constant for compound I formation from the reaction of native enzyme with H2O2 is 2.6 x 10(7) M-1 s-1. Compound I undergoes a one-electron reduction to compound II in the presence of TMB, and the rate constant for this reaction was determined to be (3.6 +/- 0.1) x 10(6) M-1 s-1. The spectral scans show that compound II accumulates in the steady state. The rate constant for compound II reduction to native enzyme by TMB obtained under steady-state conditions is (9.4 +/- 0.6) x 10(5) M-1 s-1. The results are applied to a new, more accurate assay for myeloperoxidase based upon the formation of the charge-transfer complex between TMB and its diimine final product.  相似文献   

3.
The kinetics of formation and dissociation of the horse metmyoglobin/fluoride complex has been investigated between pH 3.4 and 11. The ionic strength dependence of the reaction has been measured at integral pH values between pH 5 and 10. Hydrofluoric acid, HF, binds to metmyoglobin with a rate constant of (4.7 +/- 0. 7) x 10(4) M-1 s-1. An apparent ionization in metmyoglobin with a pKa of 4.4 +/- 0.5 influences the rate of HF binding and is attributed to the distal histidine, His-64. Protonation of His-64 increases the HF binding rate by a factor of 2.6. The fluoride anion, F-, binds to metmyoglobin with a rate constant of (5.6 +/- 1.4) x 10(-2) M-1 s-1, about 10(6) times slower than HF. Binding of either HF or F- to hydroxymetmyoglobin cannot be detected. Protonation of the distal histidine facilitates HF dissociation from the metmyoglobin/fluoride complex. HF dissociates with a rate constant of 1.9 +/- 0.3 s-1. The fluoride anion dissociates 2000 times more slowly, with a rate constant of (8.7 +/- 1.6) x 10(-4) s-1. The apparent pKa for His-64 ionization in the fluorometmyoglobin complex is 5.7 +/- 0.1. The association and dissociation rate constants are relatively independent of ionic strength with secondary kinetic salt effects sufficient to account for the ionic strength variation of both, consistent with the idea that association and dissociation of neutral HF dominate the kinetics of fluoride binding to metmyoglobin.  相似文献   

4.
Kinetics of peroxynitrite reaction with amino acids and human serum albumin   总被引:1,自引:0,他引:1  
An initial rate approach was used to study the reaction of peroxynitrite with human serum albumin (HSA) through stopped-flow spectrophotometry. At pH 7.4 and 37 degreesC, the second order rate constant for peroxynitrite reaction with HSA was 9.7 +/- 1.1 x 10(3) M-1 s-1. The rate constants for sulfhydryl-blocked HSA and for the single sulfhydryl were 5.9 +/- 0.3 and 3.8 +/- 0.8 x 10(3) M-1 s-1, respectively. The corresponding values for bovine serum albumin were also determined. The reactivity of sulfhydryl-blocked HSA increased at acidic pH, whereas plots of the rate constant with the sulfhydryl versus pH were bell-shaped. The kinetics of peroxynitrite reaction with all free L-amino acids were determined under pseudo-first order conditions. The most reactive amino acids were cysteine, methionine, and tryptophan. Histidine, leucine, and phenylalanine (and by extension tyrosine) did not affect peroxynitrite decay rate, whereas for the remaining amino acids plots of kobs versus concentration were hyperbolic. The sum of the contributions of the constituent amino acids of the protein to HSA reactivity was comparable to the experimentally determined rate constant, where cysteine and methionine (seven residues in 585) accounted for an estimated 65% of the reactivity. Nitration of aromatic amino acids occurred in HSA following peroxynitrite reaction, with nitration of sulfhydryl-blocked HSA 2-fold higher than native HSA. Carbon dioxide accelerated peroxynitrite decomposition, enhanced aromatic amino acid nitration, and partially inhibited sulfhydryl oxidation of HSA. Nitration in the presence of carbon dioxide increased when the sulfhydryl was blocked. Thus, cysteine 34 was a preferential target of peroxynitrite both in the presence and in the absence of carbon dioxide.  相似文献   

5.
The effect of free fatty acids (FFA) and non-enzymatic glycation on the binding kinetics of dansylsarcosine (DS) to human serum albumin (HSA) was studied using the stopped-flow technique. The influence of FFA on the binding parameters of 25% glycated HSA depended on the type of fatty acid. The addition of stearic, oleic and linoleic acids in a concentration of 0.3 mmol/l showed no inhibitory effects on the association rate constant (k2) value for DS binding to 25% glycated HSA (k2 without FFA: 385 +/- 10 s-1, k2 with FFA > or = 385 +/- 10 s-1). In contrast, shorter chain fatty acids (hexanoic, octanoic, decanoic, lauric and myristic acids) showed marked inhibitory effects for 0.3 mmol/l FFA (k2 range: 233 +/- 32 to 69 +/- 5 s-1) and for 0.6 mmol/l FFA (k2 range: 125 +/- 3 to 20 +/- 4 s-1). The association rate constant (k2) as well as the affinity constant (KA) of DS were markedly affected by glycation: k2 was 686 +/- 61 s-1 for 7% glycated HSA, 385 +/- 10 s-1 for 25% glycated HSA and 209 +/- 12 s-1 for 50% glycated HSA. KA decreased from 6.1 +/- 2.9 x 10(5) M-1 for 7% glycated HSA, to 5.1 +/- 0.1 x 10(5) M-1 for 25% glycated HSA and to 1.3 +/- 0.6 x 10(5) M-1 for 50% glycated HSA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Glutathione peroxidases and their mimics, e.g., ebselen or diaryl tellurides, efficiently reduce peroxynitrite/peroxynitrous acid (ONOO-/ONOOH) to nitrite and protect against oxidation and nitration reactions. Here, we report the second-order rate constant for the reaction of the reduced form of glutathione peroxidase (GPx) with peroxynitrite as (8.0 +/- 0.8) x 10(6) M-1 s-1 (per GPx tetramer) at pH 7.4 and 25 degreesC. The rate constant for oxidized GPx is about 10 times lower, (0.7 +/- 0.2) x 10(6) M-1 s-1. On a selenium basis, the rate constant for reduced GPx is similar to that obtained previously for ebselen. The data support the conclusion that GPx can exhibit a biological function by acting as a peroxynitrite reductase.  相似文献   

7.
The kinetic properties of the ba3 oxidase from Thermus thermophilus were investigated by stopped-flow spectroscopy in the temperature range of 5-70 degrees C. Peculiar behavior in the reaction with physiological substrates and classical ligands (CO and CN-) was observed. In the O2 reaction, the decay of the F intermediate is significantly slower (k' = 100 s-1 at 5 degrees C) than in the mitochondrial enzyme, with an activation energy E of 10.1 +/- 0.9 kcal mol-1. The cyanide-inhibited ba3 oxidizes cyt c522 quickly (k approximately 5 x 10(6) M-1 s-1 at 25 degrees C) and selectively, with an activation energy E of 10.9 +/- 0.9 kcal mol-1, but slowly oxidizes ruthenium hexamine, a fast electron donor for the mitochondrial enzyme. Cyt c552 oxidase activity is enhanced up to 60 degrees C and is maximal at extremely low ionic strengths, excluding formation of a high-affinity cyt c522-ba3 electrostatic complex. The thermophilic oxidase is less sensitive to cyanide inhibition, although cyanide binding under turnover is much quicker (seconds) than in the fully oxidized state (days). Finally, the affinity of reduced ba3 for CO at 20 degrees C (Keq = 1 x 10(5) M-1) was found to be smaller than that of beef heart aa3 (Keq = 4 x 10(6) M-1), partly because of an unusually fast, strongly temperature-dependent CO dissociation from cyt a32+ of ba3 (k' = 0.8 s-1 vs k' = 0.02 s-1 for beef heart aa3 at 20 degrees C). The relevance of these results to adaptation of respiratory activity to high temperatures and low environmental O2 tensions is discussed.  相似文献   

8.
We have prepared a series of biotinylated analogs of omega-conotoxin (omega CgTx) as potent, selective markers for N-type calcium channels. At pH 9.5, reaction of omega CgTx with amidocaproylbiotin succinimidyl ester gives three biotinylated conjugates, labeled at lysines 2 or 24, or at both positions. Kinetic competition assays of 125I-omega CgTx binding to rat brain synaptic membranes show that each conjugate has a similar rate constant for association (1-1.3 x 10(6) M-1 s-1) but not dissociation (1-4 x 10(-4) s-1). Comparison with rate constants obtained for the association (1.2 x 10(7) M-1 s-1) and dissociation (5 x 10(-5) s-1) of native omega CgTx indicates that while biotinylation reduces omega CgTx potency (Kdkin = k-2/k2 = 4 pM for omega CgTx), binding of these labels to membranes is nevertheless of very high affinity (Kdkin 0.1-0.3 nM).  相似文献   

9.
Fe(III) uptake by the iron-delivery and iron-scavenging protein, hen ovotransferrin has been investigated in vitro between pH 6.5 and 9. In the absence of any ferric chelate, apo-ovotransferrin loses two protons with K1a = 50 +/- 1 nM and K2a = 4.0 +/- 0.1 nM. These acid-base equilibria are independent of the interaction of the protein with bicarbonate. The interaction with bicarbonate occurs with two different affinity constants, KC = 9.95 +/- 0.15 mM and KN = 110 +/- 10 mM. FeNAc3 exchanges its Fe(III) with the C-site of the protein in interaction with bicarbonate, direct rate constants k1 = 650 +/- 25 M-1 s-1, reverse rate constant k-1 = (6.0 +/- 0.1) x 10(3) M-1 s-1 and equilibrium constant K1 = 0.11 +/- 0.01. This iron-protein intermediate loses then a single proton, K3a = 3.50 +/- 0.35 nM, and undergoes a first change in conformation followed by a two or three proton loss, first order rate constant k2 = 0.30 +/- 0.01 s-1. This induces a new modification in conformation followed by the loss of one or two protons, first order rate constant k3 = (1.50 +/- 0.05) x 10(-2) s-1. These modifications in the monoferric protein conformation are essential for iron uptake by the N-site of the protein. In the last step, the monoferric and diferric proteins attain their final state of equilibrium in about 15,000 s. The overall mechanism of iron uptake by ovotransferrin is similar but not identical to those of serum transferrin and lactoferrin. The rates involved are, however, closer to lactoferrin than serum transferrin, whereas the affinities for Fe(III) are lower than those of serum transferrin and lactoferrin. Does this imply that the metabolic function transferrins is more related to kinetics than to thermodynamics?  相似文献   

10.
We have used the pulse radiolysis technique to reduce with solvated electrons (see article) a single Fe(III) site in methemoglobin and metmyoglobin. The reduction process was followed spectrophotometrically and the reactions rate constants were measured: (see article) =6.5 +/- 1-10(10) M-1-S-1. (see article)=2.5 +/- 0.3-10(10) M-1-S-1. Approx. 60% of the (see article) have reacted with the hemin group, and the rest of the (see article) have probably reacted with the globin moiety. We followed the reaction of the reduced proteins to yield the oxyderivatives and measured the rate constants of the oxygenation process k reduced methemoglobin + O2 = 2.6 +/- 0.6-10(7) M-1-S-1 and k myoglobin + O2 = 1.8 +/- 0.2-10(7) M-1-S-1, all the rate constants were measured at pH = 6.8, I = 0.004, T = 22 +/- 2 degrees C. The high rate constant for reduced methemoglobin indicates that one-site-reduced methemoglobin is probably in the R state, as predicted for methemoglobin from X-ray analysis. The spectra of the reduced and oxygenated species were measured under similar conditions at gamma = 450-650 nm. We were able to follow slight changes in the micro-second time scale, these changes were attributed to conformational changes. We were not able to detect any reaction between the radical (see article) and the hemin group (which would result in a complex such as heme O-2). This may be due to kinetic reasons.  相似文献   

11.
The objective of this work is to reexamine the competitive degradation of deoxyribose by hydroxyl radicals (.OH) produced by the reaction between H2O2 and Fe(2+)-EDTA. The .OH radicals produced attack deoxyribose (D, rate constant kD) and eventually an .OH scavenger (S, rate constant kS). First, we examined the effect of [D], [H2O2], [Fe(2+)-EDTA], [EDTA]/[Fe2+] ratio and reaction time on the rate of D degradation, measured as the absorbance of the chromogen formed between the product of the reaction D + .OH (malondialdehyde) and thiobarbituric acid. In particular, it was showed that under our experimental conditions ([D] = 3 mM, [H2O2] = 0.85 mM, [Fe2+] = 0.13 mM), the rate of overall process is first order in Fe2+, zero order in H2O2 and is maximal for a ratio [EDTA]/[Fe2+] = 1.1. Second, the kinetics of .OH radical reaction in competition experiments between D and S (mannitol) was investigated. The results show that the ratio of the rates of D degradation in the absence (VD) and in the presence (VDS) of S should be represented by VD/VDS = 1 + ks[S]/(kD[D] + kx) where kx accounts for the rate of .OH reactions with other reagents such as Fe(2+)-EDTA, H2O2 etc . . . After having determined kx for each set of experimental conditions, we obtained the values of kS/kD by determining the variations of VD/VDS as a function of [S] and [D]. By taking kD = 1.9 x 10(9) M-1s-1 a value of kS = 1.9 x 10(9) M-1s-1 was obtained, very close to that obtained by pulse radiolysis. Finally, the validity of the established relation was confirmed for other biomolecules (methionine, k = 5.6 x 10(9)M-1s-1 and alanine, k = 3.3 x 10(8) M-1s-1). By contrast, it was not applicable to cysteine, thiourea and mercaptoethanol which was attributed to an interaction of the latter scavengers with Fe2+ and/or H2O2.  相似文献   

12.
The reduction of flavin in NADH--adrenodoxin reductase by the hydrated electron (eaq-) was investigated by pulse radiolysis. The eaq- reduced directly the flavin of the reductase to form a blue semiquinone of the enzyme. Subsequently, the semiquinone decayed by dismutation to form the oxidized and fully reduced forms of the enzyme with a second-order rate constant of 4.4 x 10(4) M-1 s-1. In the presence of equimolar NADP+, the decay of eaq- accompanied an absorption increase at 400 nm, the spectrum of which, formed transiently, is identical to that of NADP radical (NADP.). Subsequently, the transient species decayed concomitantly with the formation of the semiquinone. The rate constant in the formation of the semiquinone was independent of the concentration of the enzyme (6.1 x 10(4) s-1 at pH 7.5). From these results, it is concluded that eaq- reacts with NADP+ bound to the enzyme to form NADP. initially, and subsequently, an electron flows from the NADP. to the flavin by an intracomplex electron transfer. A similar result was obtained in the reaction of CO2- or N-methylnicotinamide radical with the NADP(+)-adrenodoxin reductase complex. These results suggest that the nicotinamide moiety of NADP+ bound to the enzyme is accessible to the solvent and masks the flavin completely.  相似文献   

13.
A ruthenium-labeled cytochrome c derivative was prepared to meet two design criteria: the ruthenium group must transfer an electron rapidly to the heme group, but not alter the interaction with cytochrome c oxidase. Site-directed mutagenesis was used to replace His39 on the backside of yeast C102T iso-1-cytochrome c with a cysteine residue, and the single sulfhydryl group was labeled with (4-bromomethyl-4' methylbipyridine) (bis-bipyridine)ruthenium(II) to form Ru-39-cytochrome c (cyt c). There is an efficient pathway for electron transfer from the ruthenium group to the heme group of Ru-39-cyt c comprising 13 covalent bonds and one hydrogen bond. Electron transfer from the excited state Ru(II*) to ferric heme c occurred with a rate constant of (6.0 +/- 2.0) x 10(5) s-1, followed by electron transfer from ferrous heme c to Ru(III) with a rate constant of (1.0 +/- 0.2) x 10(6) s-1. Laser excitation of a complex between Ru-39-cyt c and beef cytochrome c oxidase in low ionic strength buffer (5 mM phosphate, pH7) resulted in electron transfer from photoreduced heme c to CuA with a rate constant of (6 +/- 2) x 10(4) s-1, followed by electron transfer from CuA to heme a with a rate constant of (1.8 +/- 0.3) x 10(4) s-1. Increasing the ionic strength to 100 mM leads to bimolecular kinetics as the complex is dissociated. The second-order rate constant is (2.5 +/- 0.4) x 10(7) M-1s-1 at 230 mM ionic strength, nearly the same as that of wild-type iso-1-cytochrome c.  相似文献   

14.
The oxidized intermediates generated upon exposure of bovine liver catalase to hydrogen peroxide (H2O2) and superoxide radical (O2-) fluxes were examined with UV-visible spectrophotometry. H2O2 and O2- were generated by means of glucose/glucose oxidase and xanthine/xanthine oxidase systems. Serial overlay of absorption spectra in the Soret (350-450 nm) and visible (450-700 nm) regions showed that three oxidized intermediates, namely Compounds I, II and III, can be observed upon exposure of catalase to enzymatically generated H2O2 and O2-. Compound I is formed during the reaction of native enzyme with H2O2 and disappears in two ways: (i) via the catalytic reaction with H2O2 to restore native catalase and (ii) via the reaction with O2- to form Compound II. At low H2O2 concentrations (< 4.8 x 10(-9) M H2O2), Compound II reverts towards the native state mainly in a direct one-step reaction, whereas at higher H2O2 concentrations the pathway of Compound II back to the native enzyme involves Compound III. Formation of the latter from Compound II and H2O2 is irreversible and the rate constant of this reaction is 6.1 +/- 0.2 x 10(4) M-1 s-1. The formation of Compound III through the direct reaction of O2- with native enzyme has also been observed. Depending on the experimental conditions, the inactivation of catalase by O2- can be due to accumulation of Compound II ("slow" inhibition) or to the formation of Compound III ("rapid" inhibition) part of which leads to a dead end product. Formation of Compound III and of this dead end product are responsible for the irreversible inactivation in presence of an excess of H2O2.  相似文献   

15.
A new ruthenium-labeled cytochrome c derivative was designed to measure the actual rate of electron transfer to the Trp-191 radical cation and the oxyferryl heme in cytochrome c peroxidase compound I {CMPI(FeIV = O,R.+)}. The H39C,C102T variant of yeast iso-1-cytochrome c was labeled at the single cysteine residue with a tris (bipyridyl)ruthenium(II) reagent to form Ru-39-Cc. This derivative has the same reactivity with CMPI as native yCc measured by stopped-flow spectroscopy, indicating that the ruthenium group does not interfere with the interaction between the two proteins. Laser excitation of the 1:1 Ru-39-Cc-CMPI complex in low ionic strength buffer (2 mM phosphate, pH 7) resulted in electron transfer from RuII* to heme c FeIII with a rate constant of 5 x 10(5) s-1, followed by electron transfer from heme c Fe II to the Trp-191 indolyl radical cation in CMPI(FeIV = O,R*+) with a rate constant of k(eta) = 2 x 10(6) s-1. A subsequent laser flash led to electron transfer from heme c to the oxyferryl heme in CMPII-(FeIV = O,R) with a rate constant of k(etb) = 5000 s-1. The location of the binding domain was determined using a series of surface charge mutants of CcP. The mutations D34N, E290N, and A193F each decreased the values of k(eta) and k(etb) by 2-4-fold, consistent with the use of the binding domain identified in the crystal structure of the yCc-CcP complex for reduction of both redox centers [Pelletier, H., & Kraut, J. (1992) Science 258, 1748-1755]. A mechanism is proposed for reduction of the oxyferryl heme in which internal electron transfer in CMPII(FeIV = O,R) leads to the regeneration of the radical cation in CMPII-(FeIII,R*+), which is then reduced by yCcII. Thus, both steps in the complete reduction of CMPI involve electron transfer from yCcII to the Trp-191 radical cation using the same binding site and pathway. Comparison of the rate constant k(eta) with theoretical predictions indicate that the electron transfer pathway identified in the crystalline yCc-CcP complex is very efficient. Stopped-flow studies indicate that native yCcII initially reduces the Trp-191 radical cation in CMPI with a second-order rate constant ka, which increases from 1.8 x 10(8) M-1 s-1 at 310 mM ionic strength to > 3 x 10(9) M-1 s-1 at ionic strengths below 100 mM. A second molecule of yCcII then reduces the oxyferryl heme in CMPII with a second-order rate constant kb which increases from 2.7 x 10(7) M-1 s-1 at 310 mM ionic strength to 2.5 x 10(8) M-1 s-1 at 160 mM ionic strength. As the ionic strength is decreased below 100 mM the rate constant for reduction of the oxyferryl heme becomes progressively slower as the reaction is limited by release of the product yCcIII from the yCcIII-CMPII complex. Both ruthenium photoreduction studies and stopped-flow studies demonstrate that the Trp-191 radical cation is the initial site of reduction in CMPI under all conditions of ionic strength.  相似文献   

16.
The bimolecular association rate constant (kon) and dissociation rate constant (koff) of the complex between fluorescein-labeled S-peptide analogues and folded S-protein are reported. This is the first kinetic study of a protein folding reaction in which most of the starting material is already folded and only a small part (one additional helix) becomes ordered; it provides a folding landscape with a small conformational entropy barrier, and one in which kinetic traps are unlikely. Refolding and unfolding are measured under identical strongly native conditions, and the reaction is found to be two-state at low reactant concentrations. The dissociation constant (Kd) of the complex and the properties of the transition state may be calculated from the rate constants without extrapolation. The folded complex is formed fast (kon = 1.8 x 10(7) M-1 s-1) and is very stable (Kd = 6 pM) at 10 degrees C, 10 mM MOPS, pH 6.7. Charge interactions stabilize the complex by 1.4 kcal mol-1. The charge effect enters in the refolding reaction: increasing the salt concentration reduces kon dramatically and has little effect on koff. Urea and GdmCl destabilize the complex by decreasing kon and increasing koff. The slopes (m-values) of plots of ln Kd vs [cosolvent] are 0.75 +/- 0.04 and 2.8 +/- 0.3 kcal mol-1 M-1 for urea and GdmCl, respectively. The ratio mon/(mon + moff) is 0.54 +/- 0.04 for urea and 0.57 +/- 0.1 for GdmCl, where mon is the m-value for kon and moff is the m-value for koff, indicating that more than half of the sites for interaction with either cosolvent are buried in the ensemble of structures present at the transition state.  相似文献   

17.
The interaction of the N-terminal cellulose-binding domain, CBDN1, from Cellulomonas fimi beta-1,4-glucanase CenC with calcium was investigated using NMR spectroscopy and calorimetry. CBDN1 binds a single calcium ion with an equilibrium association constant of approximately 10(5) M-1 at 35 degreesC and pH 6.0. Binding is exothermic (-42 +/- 2 kJ mol-1) under these conditions and is accompanied by a small negative change in heat capacity (DeltaCp = -0.41 +/- 0.16 kJ mol-1 K-1). From an NMR line shape analysis, the rate constants for calcium association and dissociation were found to be (5 +/- 2) x 10(7) s-1 M-1 and (4.5 +/- 0.6) x 10(2) s-1, respectively. The rapid association kinetics indicate that the calcium-binding site on CBDN1 is accessible and, to the first approximation, preformed. Based on patterns of chemical shift perturbations, and structural comparisons with the Bacillus sp. 1, 3-1,4-beta-glucanases, the backbone carbonyl oxygens of Thr8, Gly30, and Asp142 and a side chain carboxyl oxygen of Asp142 are postulated to form the calcium-binding site of CBDN1. Consistent with the calcium-independent affinity of CBDN1 for cellopentaose, this exposed site is located on the face of CBDN1 opposite to that forming the oligosaccharide-binding cleft. The midpoint denaturation temperature of CBDN1 is increased by approximately 8 degreesC at pH 6.0 in the presence of saturating amounts of calcium, confirming that metal ion binding is thermodynamically linked to native-state stability.  相似文献   

18.
The aim of this work was to elucidate the effect of methanesulfonyl fluoride on the detection reaction for the determination of cholinesterase activity. The effect of methanesulfonyl fluoride was studied by monitoring the time course of the appearance and disappearance of the detection reaction product using a stopped-flow technique. The obtained experimental data were analyzed by progress curve analysis. It was found that the effect of methanesulfonyl fluoride results from a reaction between methanesulfonyl fluoride and the detection reaction product with the following rate constant: k1 = 5.33 x 10(-2) M-1s-1. In evaluating the effect of methanesulfonyl fluoride, the decomposition of this agent in water was included in the analysis because the decomposition considerably affected the reaction between methanesulfonyl fluoride and the product. The corresponding rate constant was found to be k2 = 4.9 x 10(-5) M-1s-1.  相似文献   

19.
The photochemical (lambda < 400 nm) decomposition of some monocyclic and polycyclic nitramines produces .NO2, which can be detected in the respective nitramine crystals at 77 K by EPR (electron paramagnetic resonance). In solutions of perdeutero-dimethylsulfoxide (DMSO-d6) the .NO2 produced by photolytic decomposition of dissolved nitramines can be spintrapped by the solvent to give a radical having the structure CD3-(SO2)-(NO.)-CD3. In this article, we examine this reaction for two nitramines: cyclotrimethylenetrinitramine (RDX) and hexanitrohexaazaisowurzitane (HNIW), which are energetic materials. The decay of the spin-adduct radical (I) follows first-order kinetics for both nitramines studied, having a rate constant (k) of congruent to 7.1 x 10(-4) s-1. The net growth in spin concentration of (1) measured from EPR spectra is fitted by a first-order rate equation taking into account the simultaneous competitive decay rate of spin adduct (I). Using the rate data and EPR spin concentration data, the ratio of free .NO2 produced per parent nitramine molecule is estimated as 1:1 for RDX and 4:1 for HNIW. Biological implications of trapping of .NO2 by dimethyl sulfoxide are discussed.  相似文献   

20.
The mechanism of the generation and the quantitative analysis of singlet oxygen (1O2) formed by the exposure of a hematoporphyrin derivative (HpD) to light was re-evaluated by electron spin resonance (ESR) combined with 2,2,6,6,-tetramethyl-4-piperidine (TMPD). The change from TMPD to 2,2,6,6,-tetramethyl-4-piperidine-N-oxide (TAN) has been reported to depend on singlet oxygen. However, we confirmed that this reagent also react with superoxide anion (O2-) and hydroxyl radicals (OH). Therefore, the reactions between TMPD and 1O2, O2- and OH were re-examined using a kinetic approach. We found that the generation of TAN was proportional to the concentration of TMPD and HpD, as well as to the duration and strength of the illumination. The generation of TAN was not inhibited by dimethyl-sulfoxide (DMSO) or superoxide dismutase (SOD). The reaction rate between TMPD and 1O2 was determined to be 5.0 x 10(-7) M min-1. The generation of 1O2 from HpD was 2.7 x 10(-7) M min-1 under our conditions. The competitive reaction observed between 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and TMPD for O2- or OH shows that TMPD reacts with both forms of active oxygen, but gave no ESR signal. The second-order reaction rate constant of TMPD between O2- and OH was calculated as 73 M-1 s-1 and 1.5 x 10(9) M-1 s-1, respectively. The photochemical generation of 1O2 from methylene blue, another sensitizer, was also demonstrated by this method. These results show that ESR signal of TAN can be used for the highly selective monitoring of 1O2.  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