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 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   

2.
The kinetics of the decomposition of H(2)O(2) catalyzed by Cu(II) has been studied by the initial-rate method in aqueous phosphate media at near physiological pH. The activity of the catalyst is increased by [Fe(CN)(6)](3-) and decreased by VO(3)(-), CrO(4)(2-) and Zn(II). Three reaction pathways are involved in the Cu(II)-H(2)O(2) reaction, the kinetic orders of the catalyst being 1 (rate constant k1), 2 (rate constant k2) and 3 (rate constant k3). The three pathways present fractional apparent orders (>1) in H(2)O(2) and base catalysis. The apparent activation energies associated to rate constants k1, k2 and k3 are 102+/-4, 65+/-8 and 61+/-5 kJ mol(-1). Free-radical chain mechanisms are proposed for the three pathways.  相似文献   

3.
The reduction kinetics of Fe(III)citrate, Fe(III)NTA, Co(III)EDTA-, U(VI)O(2) (2+), Cr(VI)O(4) (2-), and Tc(VII)O(4) (-) were studied in cultures of dissimilatory metal reducing bacteria (DMRB): Shewanella alga strain BrY, Shewanella putrefaciens strain CN32, Shewanella oneidensis strain MR-1, and Geobacter metallireducens strain GS-15. Reduction rates were metal specific with the following rate trend: Fe(III)citrate > or = Fe(III)NTA > Co(III)EDTA- > UO(2)(2+) > CrO(4)(2-) > TcO(4)(-), except for CrO(4) (2-) when H(2) was used as electron donor. The metal reduction rates were also electron donor dependent with faster rates observed for H(2) than lactate- for all Shewanella species despite higher initial lactate (10 mM) than H2 (0.48 mM). The bioreduction of CrO(4) (2-) was anomalously slower compared to the other metals with H(2) as an electron donor relative to lactate and reduction ceased before all the CrO(4)(2-) had been reduced. Transmission electron microscopic (TEM) and energy-dispersive spectroscopic (EDS) analyses performed on selected solids at experiment termination found precipitates of reduced U and Tc in association with the outer cell membrane and in the periplasm of the bacteria. The kinetic rates of metal reduction were correlated with the precipitation of reduced metal phases and their causal relationship discussed. The experimental rate data were well described by a Monod kinetic expression with respect to the electron acceptor for all metals except CrO(4)(2-), for which the Monod model had to be modified to account for incomplete reduction. However, the Monod models became statistically over-parameterized, resulting in large uncertainties of their parameters. A first-order approximation to the Monod model also effectively described the experimental results, but the rate coefficients exhibited far less uncertainty. The more precise rate coefficients of the first-order model provided a better means than the Monod parameters, to quantitatively compare the reduction rates between metals, electron donors, and DMRB species.  相似文献   

4.
The reaction of c-cytochromes with iron hexacyanides is similar in mechanism to the interaction of cytochromes with their physiological oxidants and reductants in that the formation of complexes precedes electron transfer. Analysis of the kinetics of oxidation and reduction of a number of c-cytochromes by solving the simultaneous differential equations defining the mechanism is possible, and allows assignment of all six rate constants describing a minimum three-step mechanism [cyto(Fe(+3)) + Fe(+2) right harpoon over left harpoon cyto (Fe(+3)) - Fe(+2) right harpoon over left harpoon cyto(Fe(+2)) - Fe(+3) right harpoon over left harpoon cyto(Fe(+2)) + Fe(+3)]. We find that the usual steady-state approximations are not valid. Furthermore, the ratio of first-order rate constants for electron transfer was approximately 1.0, and no correlation was found between any of the six rate constants and the differences in oxidation-reduction potential of the iron-hexacyanides and different cytochromes c. However, it was found that the ratio of the rate constants for complex formation between ferricytochrome c and potassium ferrocyanide and ferrocytochrome c and potassium ferricyanide was proportional to the difference in oxidation-reduction potentials. Thus the minimum three-step mechanism given above accurately describes the observed kinetic data. However, this mechanism leads to a number of conceptual difficulties. Specifically, the mechanism requires that the collision complexes formed [cyto(Fe(+3)) - Fe(CN)(6) (-4) and cyto(Fe(+2)) - Fe(CN)(6) (-3)] have very different equilibrium constants, and further requires that formation of the collision complexes be accompanied by "chemistry" to make the intermediates isoenergetic. A more complex five-step mechanism which requires that the reactants [Fe(CN)(6) (-4) and ferricytochrome c or Fe(CN)(6) (-3) and ferrocytochrome c] form a collision complex followed by a first-order process before electron transfer, was found to yield results similar to those of the three-step mechanism. However, describing the formation of the collision complex in terms of a rapid equilibrium circumvents conceptual difficulties and leads to a physically reasonable mechanism. In this mechanism the reactants are in rapid equilibrium with the collision complexes and the rate constants for complex formation are controlled by diffusion and accessibility. The collision complexes then rearrange, possibly through conformational changes and/or solvent reorganization, to yield isoenergetic intermediates that can undergo rapid reversible electron transfer. The five-step mechanism can be described by the same rate constants obtained from the three-step mechanism with the appropriate adjustments to account for rapid equilibrium. This more complex analysis associates the oxidation-reduction potential of a particular cytochrome with the relative magnitude of the first-order conversion of the oxidant and reductant collision complexes to their respective intermediates. Thus the cytochromes c control their oxidation-reduction potential by chemical and/or structural alterations. This mechanism appears to be general in that it is consistent with the observed kinetics of 11 different cytochromes c from a wide variety of sources with a range of oxidation-reduction potentials.  相似文献   

5.
The inhibitory effects of nitrate (NO3-) and nitrite (NO2-) on dissimilatory iron (FE3+) reduction were examined in a series of electron acceptor competition experiments using Shewanella putrefaciens 200 as a model iron-reducing microorganism. S. putrefaciens 200 was found to express low-rate nitrate reductase, nitrite reductase, and ferrireductase activity after growth under highly aerobic conditions and greatly elevated rates of each reductase activity after growth under microaerobic conditions. The effects of NO3- and NO2- on the Fe3+ reduction activity of both aerobically and microaerobically grown cells appeared to follow a consistent pattern; in the presence of Fe3+ and either NO3- or NO2-, dissimilatory Fe3+ and nitrogen oxide reduction occurred simultaneously. Nitrogen oxide reduction was not affected by the presence of Fe3+, suggesting that S. putrefaciens 200 expressed a set of at least three physiologically distinct terminal reductases that served as electron donors to NO3-, NO2-, and Fe3+. However, Fe3+ reduction was partially inhibited by the presence of either NO3- or NO2-. An in situ ferrozine assay was used to distinguish the biological and chemical components of the observed inhibitory effects. Rate data indicated that neither NO3- nor NO2- acted as a chemical oxidant of bacterially produced Fe2+. In addition, the decrease in Fe3+ reduction activity observed in the presence of both NO3- and NO2- was identical to the decrease observed in the presence of NO2- alone. These results suggest that bacterially produced NO2- is responsible for inhibiting electron transport to Fe3+.  相似文献   

6.
Klimacek M  Nidetzky B 《Biochemistry》2002,41(31):10158-10165
Mannitol dehydrogenases (MDH) are a family of Zn(2+)-independent long-chain alcohol dehydrogenases that catalyze the regiospecific NAD(+)-dependent oxidation of a secondary alcohol group in polyol substrates. pH and primary deuterium kinetic isotope effects on kinetic parameters for reaction of recombinant MDH from Pseudomonas fluorescens with D-mannitol have been measured in H(2)O and D(2)O at 25 degrees C and used to determine the relative timing of C-H and O-H bond cleavage steps during alcohol conversion. The enzymatic rates decreased at low pH; apparent pK values for log(k(cat)/K(mannitol)) and log k(cat) were 9.2 and 7.7 in H(2)O, respectively, and both were shifted by +0.4 pH units in D(2)O. Proton inventory plots for k(cat) and k(cat)/K(mannitol) were determined at pL 10.0 using protio or deuterio alcohol and were linear at the 95% confidence level. They revealed the independence of primary deuterium isotope effects on the atom fraction of deuterium in a mixed H(2)O-D(2)O solvent and yielded single-site transition-state fractionation factors of 0.43 +/- 0.05 and 0.47 +/- 0.01 for k(cat)/K(mannitol) and k(cat), respectively. (D)(k(cat)/K(mannitol)) was constant (1.80 +/- 0.20) in the pH range 6.0-9.5 and decreased at high pH to a limiting value of approximately 1. Measurement of (D)(k(cat)/K(fructose)) at pH 10.0 and 10.5 using NADH deuterium-labeled in the 4-pro-S position gave a value of 0.83, the equilibrium isotope effect on carbonyl group reduction. A mechanism of D-mannitol oxidation by MDH is supported by the data in which the partly rate-limiting transition state of hydride transfer is stabilized by a single solvation catalytic proton bridge. The chemical reaction involves a pH-dependent internal equilibrium which takes place prior to C-H bond cleavage and in which proton transfer from the reactive OH to the enzyme catalytic base may occur. Loss of a proton from the enzyme at high pH irreversibly locks the ternary complex with either alcohol or alkoxide bound in a conformation committed of undergoing NAD(+) reduction at a rate about 2.3-fold slower than the corresponding reaction rate of the protonated complex. Transient kinetic studies for D-mannitol oxidation at pH(D) 10.0 showed that the solvent isotope effect on steady-state turnover originates from a net rate constant of NADH release that is approximately 85% rate-limiting for k(cat) and 2-fold smaller in D(2)O than in H(2)O.  相似文献   

7.
A genetic approach was used to study (dissimilatory) ferric iron (Fe3+) reduction in Shewanella putrefaciens 200. Chemical mutagenesis procedures and two rapid plate assays were developed to facilitate the screening of Fe3+ reduction-deficient mutants. Sixty-two putative Fe3+ reduction-deficient mutants were identified, and each was subsequently tested for its ability to grow anaerobically on various compounds as sole terminal electron acceptors, including Fe3+, nitrate (NO3-), nitrite (NO2-), manganese oxide (Mn4+), sulfite (SO3(2-)), thiosulfate (S2O3(2-)), trimethylamine N-oxide, and fumarate. A broad spectrum of mutants deficient in anaerobic growth on one or more electron acceptors was identified. Nine of the 62 mutants (designated Fer mutants) were deficient only in anaerobic growth on Fe3+ and retained the ability to grow on all other electron acceptors. These results suggest that S. putrefaciens expresses at least one terminal Fe3+ reductase that is distinct from other terminal reductases coupled to anaerobic growth. The nine Fer mutants were conjugally mated with an S. putrefaciens genomic library harbored in Escherichia coli S17-1. Complemented S. putrefaciens transconjugants were identified by the acquired ability to grow anaerobically on Fe3+ as the sole terminal electron acceptor. All recombinant cosmids that conferred the Fer+ phenotype appeared to carry a common internal region.  相似文献   

8.
Batch metal sorption studies were conducted to compare the behavior of Gram-positive Bacillus subtilis and Gram-negative Escherichia coli as sorbents of Cd 2+ and Pb 2+ . A pH range from 3.0 to 6.5 was investigated at total metal concentrations of 1 2 10 -4.0 and 3.2 2 10 -5 M. Concentration apparent equilibrium sorption constants (K s n M ) and sorption capacity (S max n ) values were determined for the bacteria by fitting experimental data to one- ( n = 1) and two-site ( n = 2) Langmuir sorption isotherms. The sorption data for each of the bacteria were described well by a one-site model (r 2 > 0.9), Cd 2+ exhibited somewhat lower sorption affinities (log K s M =- 1.5 for B. subtilis , and -0.7 for E. coli ) than Pb 2+ (log K s M =-0.6 for B. subtilis and -0.8 for E. coli ). Corresponding S max values for Cd 2+ and Pb 2+ on B. subtilis were 0.36 mmole/g and 0.27 mmole/g, respectively. For E. coli Cd 2+ and Pb 2+ S max values were lower at 0.10 mmole/g and 0.21 mmole/g. A two-site sorption model yielded an improved fit for only the E. coli data with several orders of magnitude difference evident between high (Cd 2+ log K s1 M = 0.9; Pb 2+ log K s1 M = 1.5) and low (Cd 2+ log K s2 M =- 1.1; Pb 2+ log K s2 M = -1.6) affinity sorption sites. In addition, allowing for the presence of low affinity sorption (i.e., S max2 ) sites further increased the total E. coli metal sorption capacity closer to that of B. subtilis . As expected, the sorption of Cd 2+ and Pb 2+ by the bacteria exhibited a strong dependence on pH with sorption edges in the range of pH 4.2 to 5.6. The results of this study show that, despite differences in cell wall structure and composition, B. subtilis and E. coli exhibit remarkably similar sorption behavior toward Cd 2+ and Pb 2+ , respectively. These similarities can be attributed to the specific chemical reactivity of acidic functional groups (e.g., carboxyl, phosphoryl) that occur in the cell walls of both bacteria.  相似文献   

9.
Iron (2+ and 3+) is believed to transfer through the three-fold channels in the ferritin shell during iron deposition and release in animal ferritins. However, the rate of iron transit in and out through these channels has not been reported. The recent synthesis of [Fe(CN)6]3-, Prussian Blue (PB) and desferrioxamine (DES) all trapped within the horse spleen ferritin (HoSF) interior makes these measurements feasible. We report the rate of Fe2+ penetrating into the ferritin interior by adding external Fe2+ to [Fe(CN)6]3- encapsulated in the HoSF interior and measuring the rate of formation of the resulting encapsulated PB. The rate at which Fe2+ reacts with [Fe(CN)6]3- in the HoSF interior is much slower than the formation of free PB in solution and is proceeded by a lag period. We assume this lag period and the difference in rate represent the transfer of Fe2+ through the HoSF protein shell. The calculated diffusion coefficient, D approximately 5.8x10(-20) m2/s corresponds to the measured lag time of 10-20 s before PB forms within the HoSF interior. The activation energy for Fe2+ transfer from the outside solution through the protein shell was determined to be 52.9 kJ/mol by conducting the reactions at 10 approximately 40 degrees C. The reaction of Fe3+ with encapsulated [Fe(CN)6]4- also readily forms PB in the HoSF interior, but the rate is faster than the corresponding Fe2+ reaction. The rate for Fe3+ transfer through the ferritin shell was confirmed by measuring the rate of the formation of Fe-DES inside HoSF and an activation energy of 58.4 kJ/mol was determined. An attempt was made to determine the rate of iron (2+ and 3+) transit out from the ferritin interior by adding excess bipyridine or DES to PB trapped within the HoSF interior. However, the reactions are slow and occur at almost identical rates for free and HoSF-encapsulated PB, indicating that the transfer of iron from the interior through the protein shell is faster than the rate-limiting step of PB dissociation. The method described in this work presents a novel way of determining the rate of transfer of iron and possibly other small molecules through the ferritin shell.  相似文献   

10.
The present study explores the suitability of chemical reaction-based and diffusion-based kinetic models for defining the biosorption of Cu(II), Cd(II) and Pb(II) by Phormidium sp.-dominated mat. The time-course data of metal sorption by the test mat significantly (r2 = 0.932-0.999) fitted to the chemical reaction-based models namely pseudo-first-order, -second-order, and the general rate law. However, these models fail to accurately describe the kinetics of metal biosorption due either to prefixed order or unjustifiable change in rate constant and reaction order with varying concentrations of metal and biomass in the solution. The diffusion-based models, namely, the intra-particle diffusion model and the external mass transfer model fitted well to the time-course metal sorption data, thus suggesting involvement of both external and intra-particle diffusion processes in sorption of test metals by mat biomass. However, the Boyd kinetic expression clearly showed that the external mass transfer is the dominant process.  相似文献   

11.
A recent study (D. C. Cooper, F. W. Picardal, A. Schimmelmann, and A. J. Coby, Appl. Environ. Microbiol. 69:3517-3525, 2003) has shown that NO(3)(-) and NO(2)(-) (NO(x)(-)) reduction by Shewanella putrefaciens 200 is inhibited in the presence of goethite. The hypothetical mechanism offered to explain this finding involved the formation of a Fe(III) (hydr)oxide coating on the cell via the surface-catalyzed, abiotic reaction between Fe(2+) and NO(2)(-). This coating could then inhibit reduction of NO(x)(-) by physically blocking transport into the cell. Although the data in the previous study were consistent with such an explanation, the hypothesis was largely speculative. In the current work, this hypothesis was tested and its environmental significance explored through a number of experiments. The inhibition of approximately 3 mM NO(3)(-) reduction was observed during reduction of a variety of Fe(III) (hydr)oxides, including goethite, hematite, and an iron-bearing, natural sediment. Inhibition of oxygen and fumarate reduction was observed following treatment of cells with Fe(2+) and NO(2)(-), demonstrating that utilization of other soluble electron acceptors could also be inhibited. Previous adsorption of Fe(2+) onto Paracoccus denitrificans inhibited NO(x)(-) reduction, showing that Fe(II) can reduce rates of soluble electron acceptor utilization by non-iron-reducing bacteria. NO(2)(-) was chemically reduced to N(2)O by goethite or cell-sorbed Fe(2+), but not at appreciable rates by aqueous Fe(2+). Transmission and scanning electron microscopy showed an electron-dense, Fe-enriched coating on cells treated with Fe(2+) and NO(2)(-). The formation and effects of such coatings underscore the complexity of the biogeochemical reactions that occur in the subsurface.  相似文献   

12.
In contrast to its lethargy at physiological pH, horse heart cytochrome c can be oxidized at room temperature by the axial inner sphere oxidant bromomalononitrile (BMN) at higher acidities. The following stoichiometry obtains: 2Fe11 c + BrCH(CN2) + H+ leads to 2FeIII c + CH2(CN)2 + Br-, and the rate law is given by: rate = k2(FeIIc)(BMN). At an ionic strength of 1.0 (KCl), second-order rate constants vary from 300 l. per mol per sec (pH 2-3) to 0(pH 9). Below pH 6 there is a noticeable increase in rate with ionic strength while there is no specific salt effect for the process. At pH 7.4 there is no influence of added salt (0.01-1.0 M) upon the slow rate of reaction. The vast changes in rate occur over a pH region (3-6) in which only very minor changes in the visible spectrum of the cytochrome are manifest. The results are interpreted in terms of a conformational isomerism of cytochrome c in which the effective redox geometry alters from a predominantly "short C" form (in which an axial position is available for substitution) at lower pH's to a predominantly "C" form (axial positions encumbered) in the physiological region. At 5 degrees, pH 7.4, both hemes of beef heart cytochrome oxidase are oxidized by the addition of BMN (k2 = 29 plus or minus 3 l. per mol per sec). However, the reaction is inhibited by potassium cyanide and the protein containing iron(II) cyt alpha along with the cyano adduct of iron(II) or iron(III) cyt alpha3 is inert. The results demonstrate cytochrome alpha3 as the site of reaction and that alpha reduces alpha3 in the process. Cytochrome oxidase does catalyze the oxidation of cytochrome c with BMN as substrate. Taken together the results provide additional support for a recent theory and they demonstrate BMN to be an efficient probe for the effective redox geometry of a hemoprotein in solution.  相似文献   

13.
The influence of pH, ionic strength of the solution, and [Fe(CN)6]4- concentration on the rate of oxidation of sperm whale, horse, and pig oxymyoglobins, which is catalyzed by ferrocyanide ions, was studied. These myoglobins have homologous spatial structures and identical redox potentials but differ by the amount of His residues located on the protein surface. The effect of the MbO2 complexing with redox-inactive Zn2+ ion on the reaction rate was also examined. At the equimolar Zn2+ concentration, zinc ions form a stable complex with His119(GH1). It was found that the kinetic behavior of horse MbO2, which lacks His12(A10) substituted for by Gln, is fully analogous to one of sperm whale MbO2, while the oxidation of pig MbO2, three histidines of which, His12, His113(G14), and His116(G17), are replaced by Gln, is strongly inhibited. The mechanism of the catalysis was shown to involve specific binding of [Fe(CN)6]4- to the protein at the His119(GH1) site, which is in accord with the large positive electrostatic potential of this site and the presence here of a cavity large enough to accommodate [Fe(CN)6]4-. The nearby His113 and His116 residiues, which are absent in pig Mb, also play a very important role in the catalysis, because their protonation (especially of the last residue) is most likely responsible for the week oxidation of bound [Fe(CN)6]4- by dissolved oxygen.  相似文献   

14.
Exogenous electron transfer mediators employed by Fe(III)-reducing bacteria are believed to govern the kinetics and equilibrium of bioreduction of Fe(III) in solid phase. In contrast to a large number of studies on humic substances and analog anthraquinone-2,6-disulfonate (AQDS), our knowledge of other potential electron shuttles involved in Fe(III) reduction is limited. The purpose of the present study was to understand the role of cystine and cysteine in reduction of iron-rich smectite (nontronite, NAu-2) by Shewanella species. A series of abiotic and biotic experiments were conducted in nongrowth media (bicarbonate buffered, pH = 7.0). Fe(II) and cysteine concentrations were monitored over the course of the bioreduction experiments with wet chemistry, and the unreduced and reduced nontronites were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The results indicated that all Shewanella species tested here were capable of reducing cystine to cysteine. Either cystine or cysteine amendments significantly stimulated the Fe(III) bioreduction rate and extent. The initial reduction rate was linearly correlated with cystine or cysteine concentration. The reduction extent (18.7–22.3%) calculated from bioreactor with cystine or cysteine was slightly lower than those with AQDS (26.3%). Mineralogical analysis demonstrated that cystine or cysteine facilitated the reaction of smectite to illite as a result of Fe(III) bioreduction. Thus, we concluded that, in our experiments, cystine and cysteine functioned as electron carrier in the smectite reduction systems, and were favorable factors influencing smectite illitization.  相似文献   

15.
The reactions of RO(2)* radicals with Fe(H(2)O)(6)(2+) were studied, R[double bond]H; CH(3); CH(2)COOH; CH(2)CN; CH(2)C(CH(3))(2)OH; CH(2)OH; CHCl(2)/CCl(3). All these processes involve the following reactions: Fe(H(2)O)(6)(2+)+RO(2)*<==>(H(2)O)(5)Fe(III)[bond]OOR(2+) K(1) approximately 250 M(-1); (H(2)O)(5)Fe(III)[bond]OOR(2+)+H(3)O(+)/H(2)O-->Fe(H(2)O)(6)(3+)+ROOH+H(2)O/OH(-); (H(2)O)(5)Fe(III)[bond]OOR(2+)+2Fe(H(2)O)(6)(2+)-->3Fe(H(2)O)(6)(3+)+ROH; 2 RO(2)*-->Products; RO(2)*+(H(2)O)(5)Fe(III)[bond]OOR(2+)-->Fe(H(2)O)(6)(2+)+products. The values of k(1) and k(3) [reaction is clearly not an elementary reaction] approach the ligand exchange rate of Fe(H(2)O)(6)(2+), i.e. these reactions follow an inner sphere mechanism and the rate determining step is the ligand exchange step. The rate of reaction is several orders of magnitude faster than that of the Fenton reaction. Surprisingly enough the K(1) values are nearly independent of the redox potential of the radical and are considerably higher than calculated from the relevant redox potentials. These results indicate that the ROO(-) ligands considerably stabilise the Fe(III) complex, this stabilisation is smaller for radicals with electron withdrawing groups which raise the redox potential of the radical but decrease the basicity of the ROO(-) ligands, two effects which seem to nearly cancel each other. Finally, the results clearly indicate that reaction (5) is relatively fast and affects the nature of the final products. The contribution of these reactions to oxidation processes involving 'Fenton-like' processes is discussed.  相似文献   

16.
Hexavalent chromium, Cr(VI), is toxic to living systems. Widespread contamination of water and soil by Cr(VI) present a serious public health problem. Chromium-resistant bacteria can reduce and detoxify Cr(VI). Twelve bacteria resistant to high concentrations of Cr(VI) were isolated from soil enrichment cultures. Environmental parameters and kinetic parameters of Cr(VI) bioreduction by one monoculture isolate, identified by 16S rRNA gene sequence as Bacillus sp. PB2, were studied. The optimal temperature for growth and Cr(VI) reduction was 35 degrees C. The isolate grew luxuriantly and substantially reduced Cr(VI) at initial pH 7.5 to 9. Maximal Cr(VI) bioreduction occurred at initial pH 8.0. Substantial Cr(VI) bioreduction was observed in salt media, but removal efficiency was inversely related to salt concentration (1-9%). Michaelis-Menten hyperbolic equation and the Lineweaver-Burk double reciprocal plot were comparatively employed to determine the k (m) and V (max) of Cr(VI) bioreduction. A k (m) of 82.5 microg mL(-1) and V (max) of 7.78 microg mL(-1) h(-1) were calculated by nonlinear regression analysis of the hyperbola curve. Linear regression analysis of the double reciprocal plot revealed k (m) and V (max) of 80.9 microg mL(-1) and 10.6 microg mL(-1) h(-1), respectively. Time course studies displayed about 90% reduction of Cr(VI) at an initial concentration of 8,000 microg L(-1) in 8 h, with an estimated t (1/2) of 4 h. Data from time course analysis of the rate of Cr(VI) bioreduction fitted zero-order model, and the kinetic constant k was calculated to be 840 microg L(-1) h(-1). The monoculture isolate, Bacillus sp. PB2, strongly reduces Cr(VI) and could be used for bioremediation of Cr(VI)-contaminated aquatic and terrestrial environments.  相似文献   

17.
A recently described direct reading assay for beta-oxidation and for succinate oxidation in intact mitochondria using [Fe(CN)6]3- as final electron acceptor [Osmundsen & Bremer (1977) Biochem. J. 164. 621--633] has been improved by using instead tetracyano-2,2-bipyridineiron(III) [Fe(CN)4(bpy)]-, which gives a 2.6 times greater absorbance change on reduction. Some physical and kinetic properties of [Fe(CN)4(bpy)]- are described. The use of exogenous cytochrome c(III) as electron acceptor was also tested; this gives the largest absorbance change, although the absolute rate of reaction is only approx. one-third of that using [Fe(CN)6]3- or [Fe(CN)4(bpy)]-.  相似文献   

18.
Shewanella putrefaciens, a gram-negative, facultative anaerobe, is active in the cycling of iron through its interaction with Fe (hydr)oxides in natural environments. Fine-grained Fe precipitates that are attached to the outer membranes of many gram-negative bacteria have most often been attributed to precipitation and growth of the mineral at the cell surface. Our study of the sorption of nonbiogenic Fe (hydr)oxides revealed, however, that large quantities of nanometer-scale ferrihydrite (hydrous ferric oxide), goethite (alpha-FeOOH), and hematite (alpha-Fe(2)O(3)) adhered to the cell surface. Attempts to separate suspensions of cells and minerals with an 80% glycerin cushion proved that the sorbed minerals were tightly attached to the bacteria. The interaction between minerals and cells resulted in the formation of mineral-cell aggregates, which increased biomass density and provided better sedimentation of mineral Fe compared to suspensions of minerals alone. Transmission electron microscopy observations of cells prepared by whole-mount, conventional embedding, and freeze-substitution methods confirmed the close association between cells and minerals and suggested that in some instances, the mineral crystals had even penetrated the outer membrane and peptidoglycan layers. Given the abundance of these mineral types in natural environments, the data suggest that not all naturally occurring cell surface-associated minerals are necessarily formed de novo on the cell wall.  相似文献   

19.
The kinetics of bimolecular decay of alpha-tocopheroxyl free radicals (T) was studied by ESR mainly in ethanol and heptanol solvents. A second-order kinetic law was observed during the whole course of reaction (-d[T]/dt = 2k[T]2) and the following rate constants were determined with good accuracy in the temperature range 281-321 K: ethanol: log(2k) = 8.2 +/- 0.5--(6.6 +/- 0.7 kcal/mol)/(2.3RT) M-1.s-1; heptanol: log(2k) = 6.1 +/- 0.4--(4.3 +/- 0.6 kcal/mol)/(2.3RT) M-1.s-1. The global rate constant clearly increases with solvent polarity.  相似文献   

20.
Quantitative kinetic models have been developed for the reaction between peroxynitrite and membrane lipids in vesicles and for transmembrane oxidation of reactants located within their inner aqueous cores. The models were used to analyze TBARS formation and oxidation of entrapped Fe(CN)(6)(4)(-) ion in egg lecithin liposomes and several artificial vesicles. The analyses indicate that permeation of the bilayers by ONOOH and NO(2)(*), a radical formed by homolysis of the ONOOH bond, is unusually rapid but that permeation by ONOO(-) and CO(3)(*)(-), a radical formed when CO(2) is present, is negligible. Bicarbonate protects the vesicles against both membrane and Fe(CN)(6)(4)(-) oxidation by rapid competitive CO(2)-catalyzed isomerization of ONOOH to NO(3)(-); this effect is partially reversed by addition of nitrite ion, which reacts with CO(3)(*)(-) to generate additional NO(2)(*). Under medium conditions mimicking the physiological milieu, a significant fraction of the oxidants escape to inflict damage upon the vesicular assemblies. Rate constants for several elementary reaction steps, including transmembrane diffusion rates for ONOOH and NO(2)(*), were estimated from the bicarbonate dependence of the oxidative reactions.  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