首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The redox reaction between cytochrome c-551 and its oxidase from the respiratory chain of pseudomonas aeruginosa was studied by rapid-mixing techniques at both pH7 and 9.1. The electron transfer in the direction of cytochrome c-551 reduction, starting with the oxidase in the reduced and CO-bound form, is monophasic, and the governing bimolecular rate constants are 1.3(+/- 0.2) x 10(7) M-1 . s-1 at pH 9.1 and 4 (+/- 1) x 10(6) M-1 . s-1 at pH 7.0. In the opposite direction, i.e. mixing the oxidized oxidase with the reduced cytochrome c-551 in the absence of O2, both a lower absorbance change and a more complex kinetic pattern were observed. With oxidized azurin instead of oxidized cytochrome c-551 the oxidation of the c haem in the CO-bound oxidase is also monophasic, and the second-order rate constant is 2 (+/- 0.7) x 10(6) M-1 . s-1 at pH 9.1. The redox potential of the c haem in the oxidase, as obtained from kinetic titrations of the completely oxidized enzyme with reduced azurin as the variable substrate, is 288 mV at pH 7.0 and 255 mV at pH 9.1. This is in contrast with the very high affinity observed in similar titrations performed with both oxidized azurin and oxidized cytochrome c-551 starting from the CO derivative of the reduced oxidase. It is concluded that: (i) azurin and cytochrome c-551 are not equally efficient in vitro as reducing substrates of the oxidase in the respiratory chain of Pseudomonas aeruginosa; (ii) CO ligation to the d1 haem in the oxidase induces a large decrease (at least 80 mV) in the redox potential of the c-haem moiety.  相似文献   

2.
The electron transfer equilibrium and kinetics between azurin from Alcaligenes faecalis and cytochrome c551 from Pseudomonas aeruginosa have been studied. The equilibrium constant K = ([Cyt(III)] . [Az(I)])/([Cyt(II)] . [Az(II))]) = 0.5 at 25 degrees C is about seven times smaller than that observed between the cytochrome c551 and the titrations confirmed a 43-mV difference between the mid-point potentials of +266 mV and +309 mV for the Alcaligenes and Pseudomonas azurins respectively. The kinetics of the reaction between Alcaligenes azurin and Pseudomonas cytochrome c551 were investigated by the temperature-jump chemical relaxation method. Only a single relaxation mode was observed throughout the range of concentrations and temperatures examined. Thus, the slow relaxation time observed in the reaction between P. aeruginosa azurin and cytochrome c551 is not observed with the Alcaligenes azurin. The simplest mechanism that can therefore be ascribed to the investigated system is: [formula: see text]. This scheme is similar to that proposed earlier for the reaction between P. aeruginosa azurin and cytochrome c551 but does not involve the conformational transition proposed for azurin. The specific rates for the electron transfer are still fast: 1.8 x 10(6) M-1 . s-1 and 3.0 x 10(6) M-1 . s-1 respectively at 25 degrees C.  相似文献   

3.
The redox potentials of many c-type cytochromes vary with pH over the physiological pH range. We have investigated the pH dependence of redox potential for the four homologous cytochromes c-551 from Pseudomonas aeruginosa, Pseudomonas stutzeri strain 221, Pseudomonas stutzeri strain 224, and Pseudomonas mendocina . The pH dependence is due to an ionizable group that ionizes with pKox in ferricytochrome c-551 but with a higher pK, pKred , in ferrocytochrome c-551. For P. aeruginosa cytochrome c-551 it has been shown that this ionizable group is one of the heme propionic acid substituents [Moore, G. R., Pettigrew , G. W., Pitt , R. C., & Williams, R. J. P. (1980) Biochim. Biophys. Acta 590, 261-271]but the values of pKox and pKred are significantly lower in this protein than in the other three cytochromes. NMR and chemical modification studies show that for the two P. stutzeri cytochromes c-551 and P. mendocina cytochrome c-551, this propionic acid substituent is again important for the pH dependence of the redox potential. However, a histidine occurring at position 47 in their sequences hydrogen bonds to the propionic acid and thereby raises its pK. In P. aeruginosa cytochrome c-551, His-47 is substituted by Arg-47. Hydrogen-bonding schemes involving His-47 and the propionic acid are proposed.  相似文献   

4.
Effects of pH on second-order rate constants, k (25 degrees C), have been determined for the [Fe(CN)6]3- and [Co(phen)3]3+ oxidations of ferrocytochrome c-551 from Pseudomonas aeruginosa, Pseudomonas stutzeri, and Azotobacter vinelandii. For each oxidant similar directional trends are observed. With [Fe(CN)6]3-, rate constants over the pH 4-9.5 range first decrease, and then increase to plateau pH approximately equal to 9 k values of 0.96.10(5), 4.4.10(5) and 1.05.10(5) M-1.s-1, respectively. With [Co(phen)3]3+, rate constants increase in two separate well-defined stages from pH 2.5-9.5 to plateau pH approximately equal to 9 k values of 1.35.10(5), 3.6.10(5) and 1.37.10(5) M-1.s-1, respectively. From these trends, and consistent with previous NMR studies, protein pKa values of 7.16, 8.00 and 6.67, respectively, for the three reduced cytochromes c-551 are assigned to the buried propionic acid at position 7 on the haem ring. Since at pH greater than 6 the trends with pH for both [Fe(CN)6]3- and [Co(phen)3]3+ are in the same direction, it is concluded that this deprotonation results in a decrease in protein reduction potential. At pH less than 6, the trends with [Co(phen)3]3+ and [Fe(CN)6]3- are in opposite directions. Well defined pKa values of 3.6, 3.80 and 3.80 for P. aeruginosa, P. stutzeri and A. vinelandii, respectively, are observed with [Co(phen)3]3+ as oxidant. Upper limits only of pKa values less than 5.0, less than 4.1 and less than 4.5, respectively, are observed with [Fe(CN)6]3- as oxidant, which may or may not be the same as those observed for [Co(phen)3]3+. These latter pKa values are assigned to carboxylate residues at or near to the binding site(s). It is noted that charged residues are invariant on the front face (incorporating the exposed haem edge) of all three cytochromes c-551, and that there are only two carboxylates. One possibility is that the locality including both carboxylates defined by residues Asp-19, Lys-21, Lys-28 and Asp/Glu-29, serves as a binding site for both 3+ and 3- oxidants.  相似文献   

5.
Human Hageman factor, a plasma proteinase zymogen, was activated in vitro under a near physiological condition (pH 7.8, ionic strength I = 0.14, 37 degrees C) by Pseudomonas aeruginosa elastase, which is a zinc-dependent tissue destructive neutral proteinase. This activation was completely inhibited by a specific inhibitor of the elastase, HONHCOCH(CH2C6H5)CO-Ala-Gly-NH2, at a concentration as low as 10 microM. In this activation Hagemen factor was cleaved, in a limited fashion, liberating two fragments with apparent molecular masses of 40 and 30 kDa, respectively. The appearance of the latter seemed to correspond chronologically to the generation of activated Hageman factor. Kinetic parameters of the enzymatic activation were kcat = 5.8 x 10(-3) s-1, Km = 4.3 x 10(-7) M and kcat/Km = 1.4 x 10(4) M-1 x s-1. This Km value is close to the plasma concentration of Hageman factor. Another zinc-dependent proteinase, P. aeruginosa alkaline proteinase, showed a negligible Hageman factor activation. In the presence of a negatively charged soluble substance, dextran sulfate (0.3-3 micrograms/ml), the activation rate by the elastase increased several fold, with the kinetic parameters of kcat = 13.9 x 10(-3) s-1, Km = 1.6 x 10(-7) M and kcat/Km = 8.5 x 10(4) M-1 x s-1. These results suggested a participation of the Hageman factor-dependent system in the inflammatory response to pseudomonal infections, due to the initiation of the system by the bacterial elastase.  相似文献   

6.
The ionic strength dependence of the electron self-exchange rate constants of cytochromes c, c551, and b5 has been analyzed in terms of a monopole-dipole formalism (van Leeuwen, J.W. 1983. Biochim. Biophys. Acta. 743:408-421). The dipole moments of the reduced and oxidized forms of Ps. aeruginosa cytochrome c551 are 190 and 210 D, respectively (calculated from the crystal structure). The projections of these on the vector from the center of mass through the exposed heme edge are 120 and 150 D. For cytochrome b5, the dipole moments calculated from the crystal structure are 500 and 460 D for the reduced and oxidized protein; the projections of these dipole moments through the exposed heme edge are -330 and -280 D. A fit of the ionic strength dependence of the electron self-exchange rate constants gives -280 (reduced) and -250 (oxidized) D for the center of mass to heme edge vector. The self-exchange rate constants extrapolated to infinite ionic strength of cytochrome c, c551, and b5 are 5.1 x 10(5), 2 x 10(7), and 3.7 x 10(5) M-1 s-1, respectively. The extension of the monopole-dipole approach to other cytochrome-cytochrome electron transfer reactions is discussed. The control of electron transfer by the size and shape of the protein is investigated using a model which accounts for the distance of the heme from each of the surface atoms of the protein. These calculations indicate that the difference between the electrostatically corrected self-exchange rate constants of cytochromes c and c551 is due only in part to the different sizes and heme exposures of the two proteins.  相似文献   

7.
R Timkovich 《Biochemistry》1986,25(5):1089-1093
Mixtures of the dissimilatory nitrite reductase cytochrome cd1 from Pseudomonas aeruginosa and potential electron-donating proteins were prepared in both fully oxidized and fully reduced states and examined by 1H NMR spectroscopy. The relatively narrower lines of the donor proteins enabled them to be clearly observed in spectra in the presence of significant amounts of the high molecular weight cd1. Mixtures of the physiological donor (Pseudomonas ferrocytochrome c-551) and ferrocytochrome cd1 showed specific line-broadening effects on the resonances of c-551 that depended on the mole ratio of c-551 to cd1. The experimental broadening fit a model in which c-551 is in intermediate or fast exchange between free solution and a complex with cd1, with an association constant for the complex in excess of 10(4) M-1. The model yields a minimum estimate for the forward bimolecular rate constant of 5 X 10(7) M-1 s-1 and suggests that the actual value may be much larger. The complexation was independent of pH in the range of 6-8, was independent of ionic strength over a salt concentration range of 20-1000 mM, and possessed a low thermal activation barrier. Mixtures of ferricytochrome c-551 and ferricytochrome cd1 showed no observable NMR perturbations, indicating that any hypothetical complex involving the oxidized forms must follow different dynamical and/or equilibrium conditions. No observable NMR perturbations existed in spectra of mixtures of cd1 and mammalian cytochrome c or Pseudomonas azurin in either oxidation state.  相似文献   

8.
A stopped-flow investigation of the electron-transfer reaction between oxidized azurin and reduced Pseudomonas aeruginosa cytochrome c-551 oxidase and between reduced azurin and oxidized Ps. aeruginosa cytochrome c-551 oxidase was performed. Electrons leave and enter the oxidase molecule via its haem c component, with the oxidation and reduction of the haem d1 occurring by internal electron transfer. The reaction mechanism in both directions is complex. In the direction of oxidase oxidation, two phases assigned on the basis of difference spectra to haem c proceed with rate constants of 3.2 X 10(5)M-1-S-1 and 2.0 X 10(4)M-1-S-1, whereas the haem d1 oxidation occurs at 0.35 +/- 0.1S-1. Addition of CO to the reduced enzyme profoundly modifies the rate of haem c oxidation, with the faster process tending towards a rate limit of 200S-1. Reduction of the oxidase was similarly complex, with a fast haem c phase tending to a rate limit of 120S-1, and a slower phase with a second-order rate of 1.5 X 10(4)M-1-S-1; the internal transfer rate in this direction was o.25 +/- 0.1S-1. These results have been applied to a kinetic model originally developed from temperature-jump studies.  相似文献   

9.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
P Rosen  I Pecht 《Biochemistry》1976,15(4):775-786
The redox reaction between cytochrome c (Cyt c) (P-551) and the blue copper protein azurin, both from Pseudomonas aeruginosa, was studied using the temperature-jump technique. Two relaxation times were observed in a mechanism assumed to involve three equilibria. The fast relaxation time (0.4 less than tau less than 8 ms) was ascribed to the electron exchange step. The slow relaxation time (tau congruent to 37 ms) was assigned to a conformational equilibrium of the reduced azurin that was coupled through the electron exchange step to a faster conformational equilibrium of the oxidized Cyt c (P551). But because the Cyt c (P551) isomerization, being very rapid, was uncoupled from the two slower equilibria, and was assumed to involve no spectral change, the amplitude of its relaxation time (tau congruent to 0.1 ms) would be zero. At 25 degrees C and pH 7.0 the rate constants for the oxidation and reduction of Cyt c (P551) by azurin were 6.1 X 10(6) and 7.8 X 10(6) M-1 s-1, respectively; for the formation and disappearance of the reactive conformational isomer of azurin they were 12 and 17 s-1, respectively. The rates for the Cyt c (P551) isomerization could only be estimated at approximately 10(4) s-1. The thermodynamic parameters of each reaction step were evaluated from the amplitudes of the relaxations and from Eyring plots of the rate constants. Measurements of the overall equilibrium constant showed it to be temperature independent (5-35 degrees C), i.e. deltaHtot = 0. This zero enthalpy change was found to be compatible with the enthalpies calculated for the individual steps. In the electron exchange equilibrium, the values of the activation enthalpies were two to three times higher than the values published for various low molecular weight reagents in their electron exchange with copper proteins, yet the rate of exchange between Cyt c (P551) and azurin was some hundreds of times faster. This was explained in terms of the measured positive or zero entropies of activation that could result from a high level of specificity between the proteins particularly in areas of complementary charges. The mechanism of electron transfer was considered as essentially an outer sphere reaction, of which the rate could be approximated by the Marcus theory.  相似文献   

11.
Stopped-flow techniques were used to investigate the kinetics of the formation of manganese peroxidase compound I (MnPI) and of the reactions of MnPI and manganese peroxidase compound II (MnPII) with p-cresol and MnII. All of the rate data were obtained from single turnover experiments under pseudo-first order conditions. In the presence of H2O2 the formation of MnPI is independent of pH over the range 3.12-8.29 with a second-order rate constant of (2.0 +/- 0.1) x 10(6) M-1 s-1. The activation energy for MnPI formation is 20 kJ mol-1. MnPI formation also occurs with organic peroxides such as peracetic acid, m-chloroperoxybenzoic acid, and p-nitroperoxybenzoic acid with second-order rate constants of 9.7 x 10(5), 9.5 x 10(4), and 5.9 x 10(4) M-1 s-1, respectively. The reactions of MnPI and MnPII with p-cresol strictly obeyed second-order kinetics. The second-order rate constant for the reaction of MnPII with p-cresol is extremely low, (9.5 +/- 0.5) M-1 s-1. Kinetic analysis of the reaction of MnII with MnPI and MnPII showed a binding interaction with the oxidized enzymes which led to saturation kinetics. The first-order dissociation rate constants for the reaction of MnII with MnPI and MnPII are (0.7 +/- 0.1) and (0.14 +/- 0.01) s-1, respectively, when the reaction is conducted in lactate buffer. Rate constants are considerably lower when the reactions are conducted in succinate buffer. Single turnover experiments confirmed that MnII serves as an obligatory substrate for MnPII and that both oxidized forms of the enzyme form productive complexes with MnII. Finally, these results suggest the alpha-hydroxy acids such as lactate facilitate the dissociation of MnIII from the enzyme.  相似文献   

12.
Binding onto cellobiohydrolase II from Trichoderma reesei of glucose, cellobiose, cellotriose, derivatized and analogous compounds, is monitored by protein-difference-absorption spectroscopy and by titration of ligand fluorescence, either at equilibrium or by the stopped-flow technique. The data complete earlier results [van Tilbeurgh, H., Pettersson, L. G., Bhikhabhai, R., De Boeck, H. and Claeyssens, M. (1985) Eur. J. Biochem. 148, 329-334] indicating an extended active center, with putative subsites ABCD. Subsite A specifically complexes with beta-D-glucosides and D-glucose; at 25 degrees C the latter influences the concomitant binding of other ligands at neighbouring sites. For several ligands this cooperative effect for binding (at 0.33 M glucose and temperature range 4-37 degrees C) was characterized by a substantial increase of the enthalpic term (delta delta H = -35 kJ mol-1). Glucose (0.33 M) decreases the association and dissociation rate parameters of 4-methylumbelliferyl beta-D-cellobioside by one order of magnitude: k+ = (3.6 +/- 0.5) x 10(-5) M-1 s-1 versus (5.1 +/- 0.1) x 10(-6) M-1 s-1 (in the absence of glucose) and k- = (1.3 +/- 0.1) s-1 versus (14.0 +/- 0.3) s-1. As deduced from substrate-specificity studies and inhibition experiments, subsite B interacts with terminal non-reducing glucopyranosyl residues of oligomeric ligands and substrates, whereas catalytic (hydrolytic) cleavage occurs between C and D. Association constants 10-100 times higher than those for cellobiose or its glycosides were observed for D-glucopyranosyl-(1----4)-beta-D-xylopyranose and cellobionolactone derivatives, suggesting 'transition-state'-type binding for these ligands at subsite C. Although subsite D can accomodate a bulky chromophoric group (MeUmb) its preference for a glucosyl residue is reflected in the lower binding enthalpy of cellotriose (-34 kJ mol-1) as compared to cellobiose (-28.3 kJ mol-1) and MeUmb(Glc)2 (-11.6 kJ mol-1). This model indicates that oligomeric ligands (substrates) interact through cooperativity of their subunits at the extended binding site of cellobiohydrolase II.  相似文献   

13.
Electron transfer between horse heart and Candida krusei cytochromes c in the free and phosvitin-bound states was examined by difference spectrum and stopped-flow methods. The difference spectra in the wavelength range of 540-560 nm demonstrated that electrons are exchangeable between the cytochromes c of the two species. The equilibrium constants of the electron transfer reaction for the free and phosvitin-bound forms, estimated from these difference spectra, were close to unity at 20 degrees C in 20 mM Tris-HCl buffer (pH 7.4). The electron transfer rate for free cytochrome c was (2-3).10(4) M-1.s-1 under the same conditions. The transfer rate for the bound form increased with increase in the binding ratio at ratios below half the maximum, and was almost constant at higher ratios up to the maximum. The maximum electron exchange rate was about 2.10(6) M-1.s-1, which is 60-70 times that for the free form at a given concentration of cytochrome c. The activation energy of the reaction for the bound cytochrome c was equal to that for the free form, being about 10 kcal/mol. The dependence of the exchange rate on temperature, cytochrome c concentration and solvent viscosity suggests that enhancement of the electron transfer rate between cytochromes c on binding to phosvitin is due to increase in the collision frequency between cytochromes c concentrated on the phosvitin molecule.  相似文献   

14.
Rates of autoxidation of NAD(P)H initiated by hydroperoxyl radical, the acid form of superoxide anion radical which was generated by xanthine/xanthine oxidase, followed a typical autoxidation kinetic equation. Second-order rate constants for the reactions of NADPH and NADH with hydroperoxyl radical were found to be 9.82 +/- 0.13 x 10(4) M-1s-1 and 9.26 +/- 0.58 x 10(4) M-1s-1 at 25 degrees C, respectively. Rates of the reactions between NAD(P)H and superoxide to give degraded products other than NAD(P)+ were also investigated.  相似文献   

15.
Stopped-flow kinetics were made of the reaction between ascorbate-reduced Pseudomonas cytochrome oxidase and potassium ferricyanide under both N2 and CO atmospheres. Under N2 three kinetic processes were observed, two being dependent on ferricyanide concentration, with second-order rate constants of 9.6 X 10(4)M-1.s-1 and 1.5 X 10(4)M-1.s-1, whereas the other was concentration-independent, with a first-order rate constant of 0.17 +/- 0.03s-1. Measurements of their kinetic difference spectra have allowed the fastest and second-fastest phases of the reaction to be assigned to direct bimolecular reactions of ferricyanide with the haem c and haem d, moieties of the enzyme respectively. Under CO, the second-order rate constant for the reaction of the haem c was, at 1.3 X 10(5)M-1.s-1, slightly enhanced over the rate in a N2 atmosphere, but the reaction velocity of the haem d1 component was greatly decreased, being apparently limited to that of the rates of CO dissociation from the molecule (0.15s-1 and 0.03s-1). The results are compared with those obtained during a previous study of the reaction of reduced Pseudomonas cytochrome oxidase with oxidized azurin.  相似文献   

16.
Myosin light chain kinase is activated by Ca2+/calmodulin. Insights into the kinetic mechanism of this activation by Ca2+/calmodulin have now been obtained using extrinsically labeled fluorescent calmodulin, a fluorescent peptide substrate, and a stopped-flow spectrophotofluorimeter. We employed spinach calmodulin labeled with the sulfhydryl-selective probe, 2-(4-maleimidoanilino)naphthalene-6-sulfonic acid, to measure changes in the fluorescence intensity of the 2-(4-maleimidoanilino)naphthalene-6-sulfonic acid-calmodulin upon binding to rabbit skeletal muscle myosin light chain kinase. The fluorescent peptide substrate KKRAARAC(sulfobenzo-furazan)SNVFS-amide was used to measure kinase activity. Our results showed that the binding interaction could be modeled as a two-step process: a bimolecular reaction with an association rate of 4.6 x 10(7) M-1 s-1 followed by an isomerization with a rate of 2.2 s-1. Phosphorylation of the peptide during stopped-flow experiments could be modeled by a two-step process with a catalytic association rate of 6.5 x 10(6) M-1 s-1 and a turnover rate of 10-20 s-1. Our results also indicated that kinase activity occurred too rapidly for the slower isomerization rate of 2.2 s-1 to be linked specifically to the activation process.  相似文献   

17.
Compared to other monomeric heme proteins and the heme peroxidases, the Glycera dibranchiata monomer hemoglobin components III and IV exhibit very slow cyanide binding kinetics. This is agreement with the previously reported behavior of component II. Similar to component II, components III and IV have been studied under pseudo-first-order conditions at pH 6.0, 7.0, 8.0, and 9.0 by using a 100-250-fold excess of potassium cyanide at each pH. At 20 degrees C with micromolar protein concentrations, kobs for component III varies between 7.08 x 10(-5) s-1 at pH 6.0 and 100-fold cyanide excess and 1.06 x 10(-2) s-1 at pH 9.0 and 250-fold cyanide excess. For component IV, the values are 2.03 x 10(-4) s-1 for 100-fold cyanide excess at pH 6.0 and 4.13 x 10(-2) s-1 for 250-fold cyanide excess at pH 9.0. In comparison to other heme proteins, our analysis shows that the bimolecular rate constant (klapp) is small. For example, at pH 7.0, it is 3.02 x 10(-1) M-1 s-1 for component III and 1.82 M-1 s-1 for component IV, compared to 400 M-1 s-1 for sperm whale metmyoglobin, 692 M-1 s-1 for soybean metleghemoglobin a, 111 M-1 s-1 for guinea pig methemoglobin, and 1.1 x 10(5) M-1 s-1 for cytochrome c peroxidase. Our results also show that the dissociation rates (k-lapp) are extremely slow and no larger than 10(-6) s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
The binding of [125I] alpha-latrotoxin to synaptosomes from the rat brain is studied. It is shown that the constant rate of toxin association with the synaptosome receptor at 37 degrees C is equal to 8.2 +/- 1.3 x 10(7) M-1.s-1, while that of synaptosomal membrane -7.6 +/- 2.7 x 10(6) M-1 s-1. Depolarization of the synaptosome membrane induced by 55 mM KCl decreases the binding rate of toxin to the receptor, the rate constant being equal to 3.9 +/- 1.5 x 10(7) m-1 s-1. The pattern of the dissociation process of the toxin-receptor complex of synaptosomes and of synaptosomal membrane is different. In the first case dissociation follows two stages with the rate constants 3.6 x 10(-3) s-1 and 1.2/10(-4) s-1, in the second case it follows one stage with the constant equalled 2.0 x 10(-5) s-1. The quantity of the toxin binding sites on synaptosomes may vary under the action of agents modifying the activity of calcium fluxes which are induced by alpha-latrotoxin. It is supposed that a decrease in the ATP level in synaptosomes as well as deenergy of the surface membrane leads to a change in the state of the alpha-latrotoxin receptor.  相似文献   

19.
The reduction of methemoglobin by cobaltocytochrome c (Cocyt c) has been measured using nine mediators of different half-reduction potentials, Em, 7. The rate increases with the increase of Em, 7 for the mediator but dropped precipitously when it becomes more positive than the Em, 7 for the methemoglobin/hemoglobin couple. The reaction is most efficient with phenzaine methosulfate, therefore it was studied in detail. The reaction is first order in the concentrations of Cocyt c and phenazine methosulfate. The average second-order rate constant for Cocyt c + phenazine methosulfate (M) k1 leads to Cocyt c+ M-. is 2.9 x 10(4) M-1 s-1 at 25 degrees C, 0.1 M phosphate pH 7.0. There is a slight negative temperature dependence of k1 at low temperature; at higher temperatures the process has deltaH not equal to approximately 27 kJ mol-1 and deltaS not equal to approxmately - 75 J mol-1 K-1. The effect of anions reflects the dependence of Em, 7 for the methemoglobin/hemoglobin couple with various anions. There is no significant effect on k1 by the addition of inositol hexakisphosphate. The variation of k1 with pH is complicated. The experimental rate constants are compared with values calculated with the theory of nonadiabatic multiphonon process of electron tunneling.  相似文献   

20.
Laser photolysis techniques have been employed to investigate the internal electron transfer (eT) reaction within Pseudomonas aeruginosa nitrite reductase (Pa-NiR). We have measured the (d1--> c) internal eT rate for the wild-type protein and a site-directed mutant (Pa-NiR H327A) which has a substitution in the d1-heme binding pocket; we found the rate of eT to be fast, keT = 2.5 x 10(4) and 3.5 x 10(4) s-1 for the wild-type and mutant Pa-NiR, respectively. We also investigated the photodissociation of CO from the fully reduced proteins and observed microsecond first-order relaxations; these imply that upon breakage of the Fe2+-CO bond, both Pa-NiR and Pa-NiR H327A populate a nonequilibrium state which decays to the ground state with a complex time course that may be described by two exponential processes (k1 = 3 x 10(4) s-1 and k2 = 0.25 x 10(4) s-1). These relaxations do not have a kinetic difference spectrum characteristic of CO recombination, and therefore we conclude that Pa-NiR undergoes structural rearrangements upon dissociation of CO. The bimolecular rate of CO rebinding is 5 times faster in Pa-NiR H327A than in the wild-type enzyme (1.1 x 10(5) M-1 s-1 compared to 2 x 10(4) M-1 s-1), indicating that this mutation in the active site alters the CO diffusion properties of the protein, probably reducing steric hindrance. CO rebinding to the wild-type mixed valence enzyme (c3+d12+) which is very slow (k = 0.25 s-1) is proposed to be rate-limited by the c --> d1 internal eT event, involving the oxidized d1-heme which has a structure characteristic of the fully oxidized and partially oxidized Pa-NiR.  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