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.
The reaction of 3,4-dihydroxyphenylpropionic acid (dihydrocaffeic acid, hydcafH3) with chromium(III) in weak acidic aqueous solutions has been shown to take place through various oxygen-bonded intermediates. The formation of the oxygen-bonded complexes upon substitution of water molecules of the chromium(III) coordination sphere takes place in at least three stages, the first of which has an observed rate constant k1(obs)=k1K0'[hydcafH3]/[H+] where K0' corresponds to the Cr(H2O)6(3+) complex dissociation equilibrium. The second and third stages are ligand concentration independent and are thus attributed to isomerisation and chelation processes. The corresponding activation parameters are DeltaH2(not equal)=78+/-3 kJmol(-1), DeltaS2(not equal)=-49+/-9 JK(-1)mol(-1), DeltaH3(not equal)=60+/-9 kJmol(-1) and DeltaS3(not equal)=-112+/-39 JK(-1)mol(-1). The kinetic results support associative mechanisms and the nature of the electronic spectra a catecholic-type of coordination at the pH and concentration range studied and reported in this paper. The associatively activated substitution processes are accompanied by proton release causing a pH decrease. At lower acid concentration oxidation of the ligand takes place with concomitant high increase in the UV and VIS absorbance.  相似文献   

2.
Kinetics for the breakdown of the trinuclear chromium acetate cluster with a series of monoprotic and diprotic amino acid ligands and with glutathione in aqueous media have been investigated spectrophotometrically at pH 3.5–5.5 and in a temperature range of 45–60 °C. Under pseudo-first-order conditions, reactions with these ligands exhibited biphasic kinetic behavior that can be accounted for by a consecutive two-step reaction, A → B → C, where A is assumed to be a forced ion pair, B an intermediate and C is the product; experimental data fit to a biexponential equation for the transformation. Rates for kshort, klong, and kobs were determined by manual extrapolation of absorbance data or curve-fitting routines; associated activation parameters for each step of the reaction were calculated using the Eyring equation. Rates for the first and second steps of the reaction are on the order of 10−4 and 10−5 s−1, respectively. The large negative values of ΔS and smaller ΔH in the first step indicate an associative step, while high positive values of ΔS in the second step indicate dissociation. To account for the results mechanistically, the results are interpreted to be a first step of ligand exchange with a pseudo-axial aqua ligand, followed by a dissociative step involving acetate or oxo ligand displacement. The dissociative step is the rate determining step, with kobs ≈ klong.The results demonstrate reaction pathways that are available to the Cr(III) metal centers that may be physiologically relevant in the ligand-rich environment of biological systems. Under general conditions Cr(III) clusters may be expected to be broken down, unless some unique biological environment stabilizes the cluster. The present study has application to the processes related to Cr(III) transport and excretion, to potential mechanisms of Cr(III) action in a biological setting, and to the pharmacokinetics of Cr(III) supplements for animal and human consumption.  相似文献   

3.
Kinetics for the breakdown of the trinuclear chromium acetate cluster, [Cr(3)O(OAc)(6)](+), with a series of monoprotic and diprotic ligands in weakly acidic aqueous media (pH approximately 4 or approximately 5) have been investigated spectrophotometrically at 40-60 degrees C. The results point to an ion-pair equilibrium as the first step followed by associative interchange mechanism forming the mononuclear product of the reaction. Pseudo-first-order rates were determined from absorbance data and associated activation parameters were calculated using the Eyring equation. Enthalpy and entropy terms of the reactions (e.g., histidine, DeltaH(double dagger) = 75 +/- 15 kJ mol(-1), DeltaS(double dagger) = -130 +/- 25 J K(-1) mol(-1); lactic acid, DeltaH(double dagger) = 66 +/- 13 kJ mol(-1), DeltaS(double dagger) = -155 +/- 30 J K(-1) mol(-1); glycine, DeltaH(double dagger) = 31 +/- 6 kJ mol(-1), DeltaS(double dagger) = -225 +/- 45 J K(-1) mol(-1)) are consistent with an associative interchange (I(a)) mechanism, and produce a linear isokinetic plot (slope = 50 degrees C). Rates and activation parameters are comparable to those of substitution reactions of the chromium(III) hexaaqua cation. Other ligands studied included malonic acid and the amino acid, aspartic acid. Observed rates are faster than water exchange rates, but typically slower than anion substitution rates, and indicate that trinuclear chromium(III) clusters are expected to be kinetically stable in neutral to slightly acidic conditions.  相似文献   

4.
In the present study, the removal of Cr(III) from aqueous solution was studied using a new chelate-resins (b-DAEG-sporopollenin and CEP-sporopollenin). Mechanisms including ion exchange, complexation and adsorption to the surface are possible in the sorption process. Adsorption analysis results obtained at various concentrations of Cr(III) showed that the adsorption pattern on the resin followed a Langmuir isotherm. Langmuir constant Gamma max and k for Cr(III) were found as 1.23, 84.84 mmol/g for b-DAEG-sporopollenin, 133.33, 10.39 mmol/g for CEP-sporopollenin at 20 +/- 1 degrees C, respectively. In addition, kinetic and thermodynamic parameters such as enthalpy (DeltaH0), free energy (DeltaG0) and entropy (DeltaS0) were calculated and these values show that adsorption of Cr(III) on b-DAEG-sporopollenin and CEP-sporopollenin was an exothermic process and the process of adsorption was favored at high temperatures. Maximum Cr(III) removal was observed near a pH of 6.  相似文献   

5.
We investigated the dissociation of single-ring heptameric GroEL (SR1) by high hydrostatic pressure in the range 0.5-3.0 kbar. The kinetics were studied as a function of temperature in the range 15-35 degrees C. The dissociation processes at each pressure and temperature showed biphasic behavior. The slower rate (k1,obs) was confirmed to be the self-dissociation of SR1 at any specific temperature at atmospheric pressure. This dissociation was pressure independent and followed concentration-dependent first-order kinetics. The self-dissociation rates followed normal Eyring plots (In k1,obs/T vs. 1/T) from which the free energy of activation (deltaG++ = 22 +/- 0.3 kcal mol(-1)), enthalpy of activation (deltaH++ = 18 +/- 0.5 kcal mol(-1)), and entropy of activation (deltaS++ = -15 +/- 1 kcal mol(-1)) were evaluated. The effect of pressure on the dissociation rates resulted in nonlinear behavior (ln k2,obs vs. pressure) at all the temperatures studied indicating that the activation volumes were pressure dependent. Activation volumes at zero pressure (V++o) and compressibility factors (beta++) for the dissociation rates at the specific temperatures were calculated. This is the first systematic study where the self-dissociation of an oligomeric chaperonin as well as its activation parameters are reported.  相似文献   

6.
AD Vogt  E Di Cera 《Biochemistry》2012,51(30):5894-5902
For almost five decades, two competing mechanisms of ligand recognition, conformational selection and induced fit, have dominated our interpretation of ligand binding in biological macromolecules. When binding-dissociation events are fast compared to conformational transitions, the rate of approach to equilibrium, k(obs), becomes diagnostic of conformational selection or induced fit based on whether it decreases or increases, respectively, with the ligand concentration, [L]. However, this simple conclusion based on the rapid equilibrium approximation is not valid in general. Here we show that conformational selection is associated with a rich repertoire of kinetic properties, with k(obs) decreasing or increasing with [L] depending on the relative magnitude of the rate of ligand dissociation, k(off), and the rate of conformational isomerization, k(r). We prove that, even for the simplest two-step mechanism of ligand binding, a decrease in k(obs) with [L] is unequivocal evidence of conformational selection, but an increase in k(obs) with [L] is not unequivocal evidence of induced fit. Ligand binding to glucokinase, thrombin, and its precursor prethrombin-2 are used as relevant examples. We conclude that conformational selection as a mechanism for a ligand binding to its target may be far more common than currently believed.  相似文献   

7.
Brooks KM  Hampel KJ 《Biochemistry》2011,50(13):2424-2433
The glmS ribozyme is a conserved riboswitch found in numerous Gram-positive bacteria and responds to the cellular concentrations of glucosamine 6-phosphate (GlcN6P). GlcN6P binding promotes site-specific self-cleavage in the 5' UTR of the glmS mRNA, resulting in downregulation of gene expression. The glmS ribozyme has previously been shown to lack strong cation specificity when the rate-limiting folding step of the cleavage reaction pathway is measured. This does not provide data regarding cation and ligand specificities of the glmS ribozyme during the rapid ligand binding chemical catalysis events. Prefolding of the ribozyme in Mg(2+)-containing buffers effectively isolates the rapid ligand binding and catalytic events (k(obs) > 60 min(-1)) from rate-limiting folding (k(obs) < 4 min(-1)). Here we employ this experimental design to assay the cations and ligand requirements for rapid ligand binding and catalysis. We show that molar concentrations of monovalent cations are also capable of inducing the formation of the native GlcN6P binding structure but are unable to promote ligand binding and catalysis rates of >4 min(-1). Our data show that the sole obligatory role for divalent cations, for which there is crystallographic evidence, is coordination of the phosphate moiety of GlcN6P in the ligand-binding pocket. In further support of this hypothesis, our data show that a nonphosphorylated analogue of GlcN6P, glucosamine, is unable to promote rapid ligand binding and catalysis in the presence of divalent cations. Folding of the ribozyme is, therefore, relatively independent of cation identity, but the rapid initiation of catalysis upon the addition of ligand is stricter.  相似文献   

8.
A lead-dependent DNAzyme with a two-step mechanism   总被引:5,自引:0,他引:5  
Brown AK  Li J  Pavot CM  Lu Y 《Biochemistry》2003,42(23):7152-7161
A detailed biochemical and mechanistic study of in vitro selected variants of 8-17 DNAzymes is presented. Even though the 8-17 DNAzyme motif has been obtained through in vitro selection under three different conditions involving 10 mM Mg(2+) (called 8-17), 0.5 mM Mg(2+)/50 mM histidine (called Mg5), or 100 microM Zn(2+) (called 17E), all variants are shown to be the most active with Pb(2+) (8-17: k(obs) approximately 0.5 min(-1); Mg5: k(obs) approximately 2 min(-1); 17E: k(obs) approximately 1 min(-1) with 200 microM Pb(2+) at pH 5.0). For the 17E variant of the 8-17 DNAzyme, the single-turnover rate constants followed the order of Pb(2+) > Zn(2+) > Mn(2+) approximately Co(2+) > Ni(2+) > Mg(2+) approximately Ca(2+) > Sr(2+) approximately Ba(2+). The catalytic rate is half-maximal at 13.5 microM Pb(2+), 0.97 mM Zn(2+), or 10.5 mM Mg(2+), suggesting that the metal-binding affinity of the DNAzymes is in the order of Pb(2+) > Zn(2+) > Mg(2+). The Pb(2+)-dependent activity increases linearly with pH and the slope of the plot of log k(obs) versus pH is approximately 1, suggesting a single deprotonation in the rate-limiting step of the reaction. Sequence variations of the DNAzyme confirm the importance of the G*T wobble pair, the two loops and the intervening stem in maintaining the active conformation of the system. While Mg(2+) and Zn(2+) catalyze only a transesterification reaction with formation of a product containing a 2',3'-cyclic phosphate, Pb(2+) catalyzes a transesterification reaction followed by hydrolysis of the 2',3'-cyclic phosphate. Although this two-step mechanism has shown to be operative in protein ribonucleases and in the leadzyme RNAzyme, it is now demonstrated for the first time that this DNAzyme may also use the same mechanism. Therefore, the two-step mechanism is observed in metalloenzymes of all classes, and this 8-17 DNAzyme provides a simple, stable, and cost-effective model system for understanding the structure of Pb(2+)-binding sites and their roles in the two-step mechanism.  相似文献   

9.
The temperature-dependences of the second-order rate constants (k) of the reactions of the catalytic site thiol groups of two cysteine peptidases papain (EC 3.4.22.2) and actinidin (EC 3.4.22.14) with a series of seven 2-pyridyl disulphide reactivity probes (R-S-S-2-Py, in which R provides variation in recognition features) were determined at pH 6.7 at temperatures in the range 4-30 degrees C by stopped-flow methodology and were used to calculate values of DeltaS++, DeltaH++ and DeltaG++. The marked changes in DeltaS++ from negative to positive in the papain reactions consequent on provision of increase in the opportunities for key non-covalent recognition interactions may implicate microsite desolvation in binding site-catalytic site signalling to provide a catalytically relevant transition state. The substantially different behaviour of actinidin including apparent masking of changes in DeltaH++ by an endothermic conformational change suggests a difference in mechanism involving kinetically significant conformational change.  相似文献   

10.
The oxidation of D-lactobionic acid by Cr(VI) yields the 2-ketoaldobionic acid and Cr(3+) as final products when a 20-times or higher excess of the aldobionic acid over Cr(VI) is used. The redox reaction takes place through a complex multistep mechanism, which involves the formation of intermediate Cr(IV) and Cr(V) species. Cr(IV) reacts with lactobionic acid much faster than Cr(V) and Cr(VI) do, and cannot be directly detected. However, the formation of CrO(2)(2+), observed by the first time for an acid saccharide/Cr(VI) system, provides indirect evidence for the intermediacy of Cr(IV) in the reaction path. Cr(VI) and the intermediate Cr(V) react with lactobionic acid at comparable rates, being the complete rate laws for the Cr(VI) and Cr(V) consumption expressed by: -d[Cr(VI)]/dt=[k(I)+k(II)[H(+)]][lactobionicacid][Cr(VI)], where k(I)=(4.1+/-0.1) x 10(-3) M(-1) s(-1) and k(II)=(2.1+/-0.1) x 10(-2) M(-2) s(-1); and -d[Cr(V)]/dt=[k(III)[H(+)]+(k(IV)+k(V)[H(+)])[lactobionicacid]] [Cr(V)], where k(III)=(1.8+/-0.1) x 10(-3) M(-1) s(-1), k(IV)=(1.1+/-0.1) x 10(-2) M(-1) s(-1) and k(V)=(1.0+/-0.1) x 10(-2) M(-2) s(-1), at 33 degrees C. The Electron Paramagnetic Resonance (EPR) spectra show that five-co-ordinate oxo-Cr(V) bischelates are formed at pH 1-5 with the aldobionic acid bound to Cr(V) through the alpha-hydroxyacid group.  相似文献   

11.
Tertiary folding of the 160-nt P4-P6 domain of the Tetrahymena group I intron RNA involves burying of substantial surface area, providing a model for the folding of other large RNA domains involved in catalysis. Stopped-flow fluorescence was used to monitor the Mg2+-induced tertiary folding of pyrene-labeled P4-P6. At 35 degrees C with [Mg2+] approximately 10 mM, P4-P6 folds on the tens of milliseconds timescale with k(obs) = 15-31 s(-1). From these values, an activation free energy deltaG(double dagger) of approximately 8-16 kcal/mol is calculated, where the large range for deltaG(double dagger) arises from uncertainty in the pre-exponential factor relating k(obs) and delta G(double dagger). The folding rates of six mutant P4-P6 RNAs were measured and found to be similar to that of the wild-type RNA, in spite of significant thermodynamic destabilization or stabilization. The ratios of the kinetic and thermodynamic free energy changes phi = delta deltaG(double dagger)/delta deltaG(o') are approximately 0, implying a folding transition state in which most of the native-state tertiary contacts are not yet formed (an early folding transition state). The k(obs) depends on the Mg2+ concentration, and the initial slope of k(obs) versus [Mg2+] suggests that only approximately 1 Mg2+ ion is bound in the rate-limiting folding step. This is consistent with an early folding transition state, because folded P4-P6 binds many Mg2+ ions. The observation of a substantial deltaG(double dagger) despite an early folding transition state suggests that a simple two-state folding diagram for Mg2+-induced P4-P6 folding is incomplete. Our kinetic data are some of the first to provide quantitative values for an activation barrier and location of a transition state for tertiary folding of an RNA domain.  相似文献   

12.
Davidson VL 《Biochemistry》2000,39(16):4924-4928
Coupled electron transfer (ET) occurs when a relatively slow nonadiabatic ET reaction is preceded by a rapid but unfavorable adiabatic reaction that is required to activate the system for ET. As a consequence of this, the observed ET rate constant (k(ET)) is an apparent value equal to the product of the true k(ET) and the equilibrium constant for the preceding reaction step. Analysis of such reactions by ET theory may yield erroneous values for the reorganizational energy (lambda), electronic coupling (H(AB)), and ET distance that are associated with the true k(ET). If the DeltaG degrees dependence of the rate of a coupled ET reaction is analyzed, an accurate value of lambda will be obtained but the experimentally determined H(AB) will be less than the true H(AB) and the ET distance will be greater than the true distance. If the temperature dependence of the rate of a coupled ET reaction is analyzed, the experimentally determined value of lambda will be greater than the true lambda. The magnitude of this apparent lambda will depend on the magnitude of DeltaH degrees for the unfavorable reaction step that precedes ET. The experimentally determined values of H(AB) and distance will be accurate if DeltaS degrees for the preceding reaction is zero. If DeltaS degrees is positive, then H(AB) will be greater than the true value and the distance will be less than the true value. If DeltaS degrees is negative, then H(AB) will be less than the true value and the distance will be greater than the true value. Data sets for coupled ET reactions have been simulated and analyzed by ET theory to illustrate these points.  相似文献   

13.
The kinetics and mechanism of formation of the tyrosyl radical and mu-(oxo)diiron(III) cluster in the R2 subunit of ribonucleotide reductase from mouse have been examined by stopped-flow absorption and freeze-quench electron paramagnetic resonance and M?ssbauer spectroscopies. The reaction comprises (1) acquisition of Fe(II) ions by the R2 apo protein, (2) activation of dioxygen at the resulting carboxylate-bridged diiron(II) cluster to form oxidized intermediate diiron species, and (3) univalent oxidation of Y177 by one of these intermediates to form the stable radical, with concomitant or subsequent formation of the adjacent mu-(oxo)diiron(III) cluster. The data establish that an oxidized diiron intermediate spectroscopically similar to the well-characterized, formally Fe(III)Fe(IV) cluster X from the reaction of the Escherichia coli R2 protein precedes the Y177 radical in the reaction sequence and is probably the Y177 oxidant. As formation of the X intermediate (1) requires transfer of an "extra" reducing equivalent to the buried diiron cluster following the addition of dioxygen and (2) is observed to be rapid relative to other steps in the reaction, the present data indicate that the transfer of this reducing equivalent is not rate-limiting for Y177 radical formation, in contrast to what was previously proposed (Schmidt, P. P., Rova, U., Katterle, B., Thelander, L., and Gr?slund, A. (1998) J. Biol. Chem. 273, 21463-21472). Indeed, the formation of X (k(obs) = 13 +/- 3 s(-1) at 5 degrees C and 0.95 mM O(2)) and the decay of the intermediate to give the Y177 radical (k(obs) = 5 +/- 2 s(-1)) are both considerably faster than the formation of the reactive Fe(II)-R2 complex from the apo protein and Fe(II)(aq) (k(obs) = 0.29 +/- 0.03 s(-1)), which is the slowest step overall. The conclusions that cluster X is an intermediate in Y177 radical formation and that transfer of the reducing equivalent is relatively facile imply that the mouse R2 and E. coli R2 reactions are mechanistically similar.  相似文献   

14.
The gene encoding for the methionyl aminopeptidase from the hyperthermophilic archaeon Pyrococcus furiosus (PfMetAP-II; EC 3.4.11.18) has been inserted into a pET 27b(+) vector and overexpressed in Escherichia coli. The new expression system resulted in a 5-fold increase in purified enzyme obtained from a 5 L fermentor growth. The as-purified PfMetAP-II enzyme, to which no exogenous metal ions or EDTA was added, was found to have 1.2 equiv of zinc and 0.1 equiv of iron present by ICP-AES analysis. This enzyme had a specific activity of 5 units/mg, a 60-fold decrease from the fully loaded Fe(II) enzymes. When an additional 2 equiv of Zn(II) was added to the as-purified PfMetAP-II, no activity could be detected. The combination of these data with previously reported whole cell studies on EcMetAP-I further supports the suggestion that the in vivo metal ion for all MetAP's is Fe(II). Both Co(II)- and Fe(II)-loaded PfMetAP-II showed similar substrate specificities to EcMetAP-I. Substrate binding was largely affected by the amino acid in the P1 position and the length of the polypeptide. The substrates MSSHRWDW and MP-p-NA showed the smallest K(m) values while the substrates MGMM and MP-p-NA provided the highest turnover. The catalytic efficiency (k(cat)/K(m)) of PfMetAP-II for MP-p-NA at 30 degrees C was 799 500 and 340 930 M(-1) s(-1) for Co(II)- and Fe(II)-loaded PfMetAP-II, respectively. Maximum catalytic activity was obtained with 1 equiv of Co(II) or Fe(II), and the dissociation constants (K(d)) for the first metal binding site were found to be 50 +/- 15 and 20 +/- 15 nM for Co(II)- and Fe(II)-substituted PfMetAP-II, respectively. Electronic absorption spectral titration of a 1 mM sample of apo-PfMetAP-II with Co(II) provided a dissociation constant of 0.35 +/- 0.02 mM for the second metal binding site, a 17500-fold increase compared to the first metal binding site. The electronic absorption data also indicated that both Co(II) ions reside in a pentacoordinate geometry. PfMetAP-II shows unique thermostability and the optimal temperature for substrate turnover was found to be approximately 85 degrees C at pH 7.5 in 25 mM Hepes and 150 mM KCl buffer. The hydrolysis of MGMM was measured in triplicate between 25 and 85 degrees C at eight substrate concentrations ranging from 2 to 20 mM. Both specific activity and K(m) values increased with increasing temperature. An Arrhenius plot was constructed from the k(cat) values and was found to be linear over the temperature range 25-85 degrees C, indicating that the rate-limiting step in PfMetAP-II peptide hydrolysis does not change as a function of temperature. Co(II)- and Fe(II)-loaded PfMetAP-II have similar activation energies (13.3 and 19.4 kJ/mol, respectively). The thermodynamic parameters calculated at 25 degrees C are as follows: DeltaG++ = 46.23 kJ/mol, DeltaH++ = 10.79 kJ/mol, and DeltaS++ = -119.72 J.mol(-1).K(-1) for Co(II)-loaded PfMetAP; DeltaG++ = 46.44 kJ/mol, DeltaH++ = 16.94 kJ/mol, and DeltaS++ = -99.67 J.mol(-1).K(-1) for Fe(II)-loaded PfMetAP. Interestingly, at higher temperatures (> 50 degrees C), Fe(II)-loaded PfMetAP-II is more active (1.4-fold at 85 degrees C) than Co(II)-loaded PfMetAP-II.  相似文献   

15.
Wu SP  Cowan JA 《Biochemistry》2003,42(19):5784-5791
ISA type proteins mediate cluster transfer to apoprotein targets. Rate constants have been determined for cluster transfer from Schizosaccharomyces pombe ISA to apo Fd. Substitution of the cysteine residues of ISA produced derivative proteins (C72A, C136A, and C138A) that were found to be at least as active in cluster transfer reactions as the native form at 25 degrees C (k(2) approximately 170 M(-1) min(-1) for native, k(2) approximately 169 M(-1) min(-1) for C72A, k(2) approximately 206 M(-1) min(-1) for C136A, and k(2) approximately 242 M(-1) min(-1) for C138A), although the yield of cluster transfer was found to be lower as a consequence of the enhanced lability of clusters in the derivative proteins. Minor variations in rate constant for the ISA Cys derivatives do not reflect any change in the affinity of binding to the apo Fd since k(2) was found to be independent of the concentration of apo Fd over the range of 1-25 microM. The pH dependence of cluster transfer rates was found to be similar for native and C136A ISA, with an observed pK(a) of 7.8 determined from the pH profiles for cluster transfer activity of each protein. The temperature dependence of the rate constant defining the cluster transfer reaction for the wild type versus this C136A ISA derivative is distinct (DeltaH* approximately 6.3 kcal mol(-1) and DeltaS* approximately -27.3 cal K(-1) mol(-1) for native and DeltaH* approximately 2.7 kcal mol(-1) and DeltaS* approximately -38.9 cal K(-1) mol(-1) for C136A ISA). Instability of the protein-bound cluster precluded a comparison with data from pH and temperature dependencies for the two other Cys derivatives. Experiments to determine the dependence of reaction rate constants on viscosity indicate cluster transfer is rate-limiting. A comparison of cross-species rate constants for cluster transfer to apo Fd targets from Homo sapiens and S. pombe demonstrated that the identity of the Fd is less critical for promoting cluster transfer from Sp ISA (at 25 degrees C, k(2) approximately 170 M(-1) min(-1) for Sp Fd and k(2) approximately 169 M(-1) min(-1) for Hs Fd). This contrasts with an earlier observation for ISU-mediated cluster assembly [Wu, S., et al. (2002) Biochemistry 41, 8876-8885], where the rates differed for Hs and Sp target Fd's, suggesting distinct binding sites for binding of holo ISA and ISU to apo Fd.  相似文献   

16.
The reductive decomposition of both SNAP and SNOCap by ascorbate in aqueous solution (in the presence of EDTA) was thoroughly investigated. Nitric oxide (NO) release from the reaction occurs in an ascorbate concentration and pH dependent manner. Rates and hence NO release increased drastically with increasing pH, signifying that the most highly ionized form of ascorbate is the more reactive species. The experiments were monitored spectrophotometrically, and second-order rate constants calculated at 37 degrees C for the reduction of SNAP are k(b)=9.81+/-1.39 x 10(-3) M(-1) s(-1) and k(c)=662+/-38 M(-1) s(-1) and for SNOCap are k(b)=2.57+/-1.29 x 10(-2) M(-1) s(-1) and k(c)=49.7+/-1.3 M(-1) s(-1). k(b) and k(c) are the second-order rate constants via the ascorbate monoanion (HA-) and dianion (A2-) pathways, respectively. Activation parameters were also calculated and are DeltaHb++ =93+/-7 kJ mol(-1), DeltaSb++ =15+/-2 J K(-1) mol(-1) and DeltaHc++ =51+/-5 kJ mol(-1), DeltaSc++ =-28+/-3 J K(-1) mol(-1) with respect to the reactions involving SNAP. Those for the reaction between SNOCap and ascorbate were calculated to be DeltaHb++ =63+/-11 kJ mol(-1), DeltaSb++ =-71+/-20 J K(-1) mol(-1) and DeltaHc++ =103+/-7 kJ mol(-1), DeltaSc++ =118+/-8 J K(-1) mol(-1). The effect of Cu2+/Cu+ ions on the reductive decompositions of these S-nitrosothiols was also investigated in absence of EDTA. SNOCap exhibits relatively high stability at near physiological conditions (37 degrees C and pH 7.55) even in the presence of micromolar concentrations of Cu2+, with decomposition rate constant being 0.011 M(-1) s(-1) in comparison to SNAP which is known to be more susceptible to catalytic decomposition by Cu2+ (second-order rate constant of 20 M(-1) s(-1) at pH 7.4 and 25 degrees C). It was also observed that the reductive decomposition of SNAP is not catalyzed by alkali metal ions, however, there was an increase in rate as the ionic strength increases from 0.2 to 0.5 mol dm(-3) NaCl.  相似文献   

17.
The mechanism of activation thioamide-pyridine anti-tuberculosis prodrugs is poorly described in the literature. It has recently been shown that ethionamide, an important component of second-line therapy for the treatment of multi-drug-resistant tuberculosis, is activated through an enzymatic electron transfer (ET) reaction. In an attempt to shed light on the activation of thioamide drugs, we have mimicked a redox process involving the thionicotinamide (thio) ligand, investigating its reactivity through coordination to the redox reversible [Fe(III/II)(CN)(5)(H(2)O)](2-/3-) metal center. The reaction of the Fe(III) complex with thionicotinamide leads to the ligand conversion to the 3-cyanopyridine species coordinated to a Fe(II) metal center. The rate constant, k(et)=10 s(-1), was determined for this intra-molecular ET reaction. A kinetic study for the cross-reaction of thionicotinamide and [Fe(CN)(6)](3-) was also carried out. The oxidation of thionicotinamide by [Fe(CN)(6)](3-) leads to formation of mainly 3-cyanopyridine and [Fe(CN)(6)](4-) with a k(et)=(5.38+/-0.03) M(-1)s(-1) at 25 degrees C, pH 12.0. The rate of this reaction is strongly dependent on pH due to an acid-base equilibrium related to the deprotonation of the R-SH functional group of the imidothiol form of thionicotinamide. The kinetic results reinforced the assignment of an intra-molecular mechanism for the ET reaction of [Fe(III)(CN)(5)(H(2)O)](2-) and the thioamide ligand. These results can be valuable for the design of new thiocarbonyl-containing drugs against resistant strains of Mycobacterium tuberculosis by a self-activating mechanism.  相似文献   

18.
Intracellular reduction of carcinogenic Cr(VI) generates Cr-DNA adducts formed through the coordination of Cr(III) to DNA phosphates (phosphotriester-type adduct). Here, we examined the role of Cr(III)-DNA adducts in mutagenesis induced by metabolism of Cr(VI) with cysteine. Reduction of Cr(VI) caused a strong oxidation of 2', 7'-dichlorofluoroscin (DCFH) and extensive Cr-DNA binding but no DNA breakage. Cr-DNA adducts induced unwinding of supercoiled plasmids and structural distortions in the DNA helix as detected by decreased ethidium bromide binding. Propagation of Cr-treated pSP189 plasmids in human fibroblasts led to a dose-dependent formation of the supF mutants and inhibition of replication. Blocking of Cr(III)-DNA binding by occupation of DNA phosphates with Mg(2+) or by sequestration of Cr(III) by inorganic phosphate or EDTA eliminated mutagenic responses and restored a normal yield of replicated plasmids. Dissociation of Cr(III) from DNA by a phosphate-based reversal procedure returned mutation frequency to background levels. The mutagenic responses at the different phases of the reduction reaction were unrelated to the amount of reduced Cr(VI) but reflected the number and the spectrum of Cr(III)-DNA adducts that were formed. Ternary cysteine-Cr(III)-DNA adducts were approximately 4-5 times more mutagenic than binary Cr(III)-DNA adducts. Although intermediate reaction products (CrV/IV, thiyl radicals) were capable of oxidizing DCFH, they were insufficiently reactive to damage DNA. Single-base substitutions at G/C pairs were the predominant type of Cr-induced mutations. The majority of mutations occurred at the sites where G had adjacent purine in the 3' or 5' position. Overall, our results present the first evidence that Cr(III)-DNA adducts play the dominant role in the mutagenicity caused by the metabolism of Cr(VI) by a biological reducing agent.  相似文献   

19.
A new chromium complex, (bis(hydroxyethyl)amino-tris(hydroxymethyl)methane)oxochromate(V), has been characterized by epr spectroscopy. The chromium(V) complex was formed by the ligand displacement reaction of bis(2-ethyl-2-hydroxybutanato) oxochromate(V) with bis(hydroxyethyl)amino-tris(hydroxy-methyl)methane (BT). Both epr and kinetic data indicate that the reaction proceeds through a chromium(V) intermediate. Kinetics of formation of the intermediate exhibit a rate saturation at higher [BT] (> 30 mM) indicating a rate law constituting an equilibrium between the parent Cr(V) complex and the bis-tris ligand followed by a pure first order process. The g-value of the intermediate is consistent with a Cr(V) complex in which the BT is coordinated in a bidentate fashion replacing a coordinated hydroxy butanoic acid ligand, affording a mixed ligand complex. The equilibrium step (K = 36 M-1) consists of monodentate coordination by the BT ligand and the limiting first order rate constant (1.9 x 10(-2) s-1) manifests the rate of chelation by the polydentate ligand. The intermediate is converted to the product upon further chelation through the complete displacement of the remaining 2-ethyl-2-hydroxy butanoic acid by a first order process (k = 0.023 s-1). The epr data support a pair of products that are in rapid equilibrium. In these products, BT functions either as a tetra or a penta-dentate ligand coordinating through four or five alkoxy sites. The enthalpy and entropy of activations related to the two chelation steps were found to be 32 +/- 2 kJ/mol and -(1.7 +/- 0.2) x 10(2) J/mol K for the intermediate, and 36 +/- 1 kJ/mol and -(1.5 +/- 0.2) x 10(2) J/mol K for the product. Our data support an associative mechanism for the chelation steps. The Cr(V)-BT product is more stable than the parent complex. The second order disproportionation rate constant for the Cr(V)-BT complex was evaluated to be 0.1 M-1 s-1 compared to 8.0 M-1 s-1 for the parent complex. This is the first example of a chromium(V) complex with a non-macrocyclic ligand coordinating through oxygen donor atoms which is stable in aqueous solution at neutral pH over a long period of time.  相似文献   

20.
Buckman J  Miller SM 《Biochemistry》2000,39(34):10521-10531
The transient kinetics of the reaction of the estrogen binding protein (EBP1) from Candida albicans in which hydride is transferred from NADPH to trans-2-hexenal (HXL) in two half-reactions were analyzed using UV-visible spectrophotometric and fluorometric stopped-flow techniques. The simplest model of the first half-reaction involves four steps including very rapid, tight binding (K(d) 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