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.
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   

2.
During the reaction of oxyhemoglobin (HbO2) with nitrite, the concentration of residual nitrite, nitrate, oxygen, and methemoglobin (Hb+) was determined successively. The results obtained at various pH values indicate the following stoichiometry for the overall reaction: 4HbO2 + 4NO2- 4H+ leads to 4Hb+ + 4NO3- + O2 + 2H2 O (Hb denotes hemoglobin monomer). NO2- binds with methemoglobin noncooperatively with a binding constant of 340 M-1 at pH 7.4 and 25 degrees C. Thus, the major part of Hb+ produced is aquomethemoglobin, not methemoglobin nitrite, when less than 2 equivalents of nitrite is used for the oxidation.  相似文献   

3.
Stimulation of the oxygen (O2) metabolism of isolated human neutrophilic leukocytes resulted in oxidation of hemoglobin of autologous erythrocytes without erythrocyte lysis. Hb oxidation could be accounted for by reduction of O2 to superoxide (O-2) by the neutrophils, dismutation of O-2 to yield hydrogen peroxide (H2O2), myeloperoxidase-catalyzed oxidation of chloride (Cl-) by H2O2 to yield hypochlorous acid (HOCl), the reaction of HOCl with endogenous ammonia (NH+4) to yield monochloramine ( NH2Cl ), and the oxidative attack of NH2Cl on erythrocytes. NH2Cl was detected when HOCl reacted with the NH+4 and other substances released into the medium by neutrophils. The amount of NH+4 released was sufficient to form the amount of NH2Cl required for the observed Hb oxidation. Oxidation was increased by adding myeloperoxidase or NH+4 to increase NH2Cl formation. Due to the volatility of NH2Cl , Hb was oxidized when neutrophils and erythrocytes were incubated separately in a closed container. Oxidation was decreased by adding catalase to eliminate H2O2, dithiothreitol to reduce HOCl and NH2Cl , or taurine to react with HOCl or NH2Cl to yield taurine monochloramine . NH2Cl was up to 50 times more effective than H2O2, HOCl, or taurine monochloramine as an oxidant for erythrocyte Hb, whereas HOCl was up to 10 times more effective than NH2Cl as a lytic agent. NH2Cl contributes to oxidation of erythrocyte components by stimulated neutrophils and may contribute to other forms of neutrophil oxidative cytotoxicity.  相似文献   

4.
Following flash excitation, oxygen pulses and fluorescence kinetics in the time range 0-16 mus were studied in the alga Chlorella pyrenoidosa during incubation with various concentrations of hydroxylamine. The obtained results could be explained considering four effects of hydroxylamine. 1. Hydroxylamine removes (reduces) oxidizing equivalents, generated in the water-splitting system by flash excitation. This process does not markedly affect the fluorescence yield kinetics between 0 and 16 mus following the ignition of a flash and reaches a constant rate within a few minutes, but possibly within a few seconds, after addition of hydroxylamine. In a sequence of flashes separated by dark time td, the steady-state oxygen yield in the flashes is exp(-ktd), the yield at td=0 being taken equal to 1, where k=(0.1 + beta[NH2OH])s-1, with [NH2OH] in mM and beta=0.6 mM-1, provided [NH2OH]greater than or equal to 0.5 mM. 2. An inhibition between Z, the physiological donor and the oxidized reaction center pigment P+ occurs, proceeding as exp (-kiti)where ti is the incubation time with hydroxylamine and ki=(alpha[NH2OH]) min-1, with [NH2OH] in mM and alpha=0.14 mM-1. This process not only inhibits oxygen evolution capability, but also decreases the amplitude of the fluorescence yield difference deltaphi=phi(16 mus)-phi(2 mus) induced by a flesh in the steady state. In a fraction of the reaction centers this inhibition occurs "immediately" after the addition of hydroxylamine. These observations, combined with the conslusion of Cheniae and Martin (1971, Plant Physiol. 47, 568-575) that the inhibition of the Hill reaction is related to the extraction of bound manganese indicate that the reaction between Z and P+ requires bound manganese. 3. In the inhibited centers a second donor for P+, D, connected to an entry site for the artificial electron donor hydroxylamine becomes apparent. 4. A flash-induced oxygen uptake signal was observed in the presence of hydroxylamine, which was shown to be caused by a system II reaction. The effects under (1) and (4) were reversed in the dark if hydroxylamine was removed by washing. The effects under (2) and (3) were reversed during illumination of a washed sample.  相似文献   

5.
The binding of cyanide to both oxidized and ascorbate-reduced forms of Pseudomonas cytochrome c-551 oxidase was investigated. Spectral studies on the oxidized enzyme and its apoprotein showed that the ligand can bind to both the c and d, haem components of the molecule, and kinetic observations indicated that both chromophores reacted, under a variety of conditions, with very similar rates. Cyanide combination velocities were dependent on ligand concentration, and increasing the pH also accelerated the reaction; the second-order rate constant was estimated as approx. 0.2M-1 . s-1 at pH 7.0. The binding of cyanide to the protein was observed to have a considerable influence on reduction of the enzyme by ascorbate. Spectral and kinetic observations have revealed that the species haem d13+-cyanide and any unbound haem c may react relatively rapidly with the reductant, but the behaviour of cyanide-bound haem c indicates that it may not be reduced without prior dissociation of the ligand, which occurs relatively slowly. The reaction of reduced Pseudomonas cytochrome oxidase with cyanide is radically different from that of the oxidized protein. In this case the ligand only binds to the haem d1 component and reacts much more rapidly. Stopped-flow kinetic measurements showed the binding to be biphasic in form. Both the rates of these processes were dependent on cyanide concentration, with the fast phase having a second-order rate constant of 9.3 X 10(5) M-1 . s-1 and the slow phase one of 2.3 X 10(5) M-1 . s-1. The relative proportions of the two phases also showed a dependency on cyanide concentration, the slower phase increasing as the cyanide concentration decreased. Computer simulations indicate that a reaction scheme originally proposed for the reaction of the enzyme with CO is capable of providing a reasonable explanation of the experimental results. Static-titration data of the reduced enzyme with with cyanide indicated that the binding was non-stoicheiometric, the ligand-binding curve being sigmoidal in shape. A Hill plot of the results yielded a Hill coefficient of 2.6.  相似文献   

6.
Lumbricus terrestris HbO2 and HbCO dissociated below pH 5.0; a time-dependent alteration to the met form occurred at pH less than 5 and pH less than 4.5, respectively. The extent of dissociation was unaffected by alkaline earth cations but was decreased by an increase in ionic strength. HbO2 and HbCO exposed to pH 4.0-4.8 were centrifuged to obtain the undissociated pellet (P1) and dissociated supernatant (S1) fractions. S1 was reassociated at pH 7.0 by dialysis against various buffers and then centrifuged to obtain the reassociated pellet (P2) and unreassociated supernatant (S2) fractions. Reassociation was possible only if S1 was dialyzed against water prior to return to neutral pH; otherwise precipitation occurred starting at about pH 5.3. The extent of reassociation varied from about 40 to 80%, was usually higher for HbCO than HbO2, and was unaffected by an increase in ionic strength or by Ca(II). Gel filtration of P2 on Sephacryl S-300 at neutral pH gave one peak IaR, eluting at a slightly greater volume than the native Hb; S1 and S2 gave in addition, three peaks, Ib (200 kDa), II (65 kDa), and III (18 kDa). Sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that P2 was slightly deficient in subunit M relative to the Hb, that Ib was deficient in subunits D1 and D2 and that II and III consisted of subunits D1 + D2 + T and subunit M, respectively. Scanning transmission electron microscopy of P2 showed that it was smaller than the native hemoglobin: 25 nm in diameter and 16 nm in height, instead of 30 X 20 nm. Comparison of the results of the dissociations of Lumbricus Hb at alkaline pH (Kapp, O. H., Polidori, G., Mainwaring, M., Crewe, A. V., Vinogradov, S. N. (1984) J. Biol. Chem. 259, 628-639) with those obtained in this study suggested that the Hb quaternary structure was not multimeric and that an alternative model had to be considered. In the proposed model it is assumed that subunits D1 and D2 form a scaffolding or "bracelet," decorated with 12 complexes of M and T subunits.  相似文献   

7.
Traces of iron, when complexed with either EDTA or diethylenetriaminepentaacetic acid (DTPA), catalyze an OH.-producing reaction between H2O2 and paraquat radical (PQ+.): H2O2 + PQ+.----PQ++ + OH. + OH-.[1]. Kinetic studies show that oxidation of formate induced by this reaction occurs by a Fenton-type mechanism, analagous to that assumed in the metal-catalyzed Haber-Weiss reaction, in which the rate determining step is H2O2 + Fe2+ (chelator)----Fe3+(chelator) + OH. + OH-,[7]; with k7 = 7 X 10(3) M-1 s-1 for EDTA and 8 X 10(2) M-1 s-1 for DTPA at pH 7.4. PQ+. rapidly reduces both Fe3+ (EDTA) and Fe3+ (DTPA), and hence allows both agents to catalyze [1] with comparable efficiency, in contrast to the much lower efficiency reported for the latter as a catalyst for the Haber-Weiss reaction. The catalytic properties of these chelating agents is attributed to their lowering of E0 (Fe3+/Fe2+) by 0.65 V, thus making [7] thermodynamically possible at pH 7. Approximately 2.5% of the OH. produced is consumed by internal or "cage" reactions, which decompose the chelator and produce CO2; however, the majority (97%) diffuses into the bulk solution and participates in competitive reactions with OH. scavengers.  相似文献   

8.
The reactions of cis-[PtCl(NH3)2(H2O)]+ with L-methionine have been studied by 1D 195Pt and 15N NMR, and by 2D[1H, 15N] NMR. When the platinum complex is in excess, the initial product, cis-[PtCl(NH3)2(Hmet-S)]+ undergoes slow ring closure to [Pt(NH3)2(Hmet-N,S)]2+. Slow ammine loss then occurs to give the isomer of [PtCl(NH3)(Hmet-N,S)]+ with chloride trans to sulfur. When methionine is in excess, a reaction sequence is proposed in which trans-[PtCl(NH3)(Hmet-S)2]+ isomerises to the cis-isomer, with subsequent ring closure reactions leading to cis-[Pt(Hmet-N,S)2]2+. Near pH 7, methionine is unreactive toward cis-[PtCl(OH)(NH3)2]. By contrast, L-cysteine reacts readily with cis-[PtCl(OH)(NH3)2] at pH 7, but there were many reaction products, including bridged species. Cis-[PtCl(OH)(NH3)2] reacts with reduced thiols in ultrafiltered plasma but these are oxidized if the plasma is not fresh or appropriately stored. With very low concentrations of the platinum complexes (35.5 microM), HPLC experiments (UV detection at 305 nm) indicate that the thiolate (probably cysteine) reactions become simpler as bridging becomes less important.  相似文献   

9.
Dismutation of dihydrofolate by dihydrofolate reductase   总被引:1,自引:0,他引:1  
R L Blakley  L Cocco 《Biochemistry》1984,23(11):2377-2383
Degradation of 7,8-dihydrofolate (H2folate) in the presence of dihydrofolate reductase (DHFR) has been shown due not to an oxygenase activity of the reductase as previously reported but to dismutation of H2folate to folate and 5,6,7,8-tetrahydrofolate (H4folate). The reaction can be followed spectrophotometrically or by analysis of the reaction mixture by high-performance liquid chromatography (HPLC). The products have also been isolated and characterized. Oxygen uptake during the reaction is much less than stoichiometric with H2folate disappearance and is attributed to autoxidation of the H4folate formed. The dismutation activity is a property of highly purified Streptococcus faecium DHFR isoenzyme 2 (but not isoenzyme 1) and of Lactobacillus casei DHFR, but not of bovine liver DHFR. The activity is dependent on tightly bound NADP+ and/or NADPH. Removal of the nucleotide results in loss of dismutation activity, which is restored by adding NADP+ or NADPH. Maximum activity is obtained when approximately 1 mol equiv of nucleotide is added per mol of DHFR. It is proposed that in the dismutation reaction bound NADP(H) is alternately reduced and oxidized by incoming molecules of H2folate with release of folate and H4folate, respectively. The relatively slow rate of folate formation presumably limits the rate of the overall reaction. The equilibrium constant for the dismutation reaction is 19.4 +/- 7.4 at 22 degrees C and pH 7.0. Calculation of standard oxidation-reduction potentials at pH 7 gave values of -0.230 V for the H2folate/H4 folate pair and -0.268 V for the folate/H2folate pair. The mechanism by which NADP+ is retained by the enzyme from some sources during purification procedures is unclear.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
We have recorded the C-2 proton resonances of the histidines of carbonmonoxyhaemoglobin A and of four abnormal human HbCOs in different buffers and at different concentrations of haemoglobin. Resonance H assigned by Perutz et al. (1985) to His HC3(146) beta, is present at both pH 7.30 and pH 6.90, but somewhat broadened when recorded in 5 to 10% HbCO A in 0.1 M-bis-Tris. The broadening disappears on tenfold dilution of the Hb with bis-Tris and the resonance then stands out sharply. Resonance H is absent at both Hb concentrations in HbCO Cowtown (His HC3(146) beta----Leu). HbCO Fort de France (His CD3(45) alpha----Arg) in 0.1 M-bis-Tris of pH 6.90 has a spectrum similar to that of HbCO A. In the same buffer a resonance marked L by Russu et al. (1982) is absent from the spectrum of Hb Abbruzzo (His H21(143) beta----Arg), whereas resonance H is present. Hb Barcelona contains an additional histidine in position FG1(94) beta; in 0.1 M-bis-Tris buffer of pH 6.90 its resonance is not resolved and resonance H is either shifted or broadened. The resonances of both histidines are resolved in phosphate buffer. At pH 6.90, spectra in 0.1 M-bis-Tris buffer are similar to those previously recorded in 0.2 M-HEPES. Addition of 0.1 M-KCl produces marked changes. Replacement of bis-Tris by 0.2 M-KCl + 0.2 M-phosphate gives rise to a different and much better resolved spectrum.  相似文献   

11.
The superoxide dismutative action of Cu-penicillamine was examined by pulse radiolysis. The second order rate constand of the reaction wpith superoxide was 0.4 +/- o.2.10(9) M-1.s-1, comparable to the action of Fe and Mn-superoxide dismutases. No marked pH-dependence was seen. Neither ethylene diamine tetraacetic acid nor cyanide affected the catalytic action of Cu-penicillamine. The cyanide resistant reactivity as well as further X-ray photoelectron spectrometric measurements supported the suggestion of a Cu(I) stabilized sulphur radical being the active species involved in the catalysis of superoxide dismutation.  相似文献   

12.
S A Ensign  M R Hyman  P W Ludden 《Biochemistry》1989,28(12):4973-4979
The inhibition of purified carbon monoxide dehydrogenase from Rhodospirillum rubrum by cyanide was investigated in both the presence and absence of CO and electron acceptor. The inhibition was a time-dependent process exhibiting pseudo-first-order kinetics under both sets of conditions. The true second-order rate constants for inhibition were 72.2 M-1 s-1 with both substrates present and 48.9 and 79.5 M-1 s-1, respectively, for the reduced and oxidized enzymes incubated with cyanide. CO partially protected the enzyme against inhibition after 25-min incubation with 100 microM KCN. Dissociation constants of 8.46 microM (KCN) and 4.70 microM (CO) were calculated for the binding of cyanide and CO to the enzyme. Cyanide inhibition was fully reversible under an atmosphere of CO after removal of unbound cyanide. N2 was unable to reverse the inhibition. The competence of nickel-deficient (apo) CO dehydrogenase to undergo activation by NiCl2 was unaffected by prior incubation with cyanide. Cyanide inhibition of holo-CO dehydrogenase was not reversed by addition of NiCl2. 14CN- remained associated with holoenzyme but not with apoenzyme through gel filtration chromatography. These findings suggest that cyanide is a slow-binding, active-site-directed, nickel-specific, reversible inhibitor of CO dehydrogenase. We propose that cyanide inhibits CO dehydrogenase by being an analogue of CO and by binding through enzyme-bound nickel.  相似文献   

13.
The reduction of flavin in hepatic NADH-cytochrome b5 reductase by the hydrated electron (eaq-) was investigated by pulse radiolysis. The eaq- reduced the flavin of NADH-cytochrome b5 reductase to form the red semiquinone between pH 5 and 9. The spectrum of the red semiquinone differs from that of enzyme reduced by dithionite in the presence of NAD+. After the first phase of the reduction, conversion of the red to blue semiquinone was observed at acidic pH. Resulting products are the blue (neutral) or red (anionic) semiquinone or a mixture of the two forms. The pK value for this flavin radical was approximately 6.3. Subsequently, the semiquinone form reacted by dismutation to form the oxidized and the fully reduced forms of the enzyme with a rate constant of 1 x 10(3) M-1 s-1 at pH 7.1. In the presence of NAD+, eaq- reacted with NAD+ to yield NAD(.). Subsequently, NAD. transferred an electron to NAD+-bound oxidized enzyme to form the blue and red semiquinone or mixture of the two forms of the enzyme, where pK value of this flavin radical was approximately 6.3. The blue semiquinone obtained at acidic pH was found to convert to the red semiquinone with a first order rate constant of 90 s-1, where the rates were not affected by pH or the concentration of NAD+. The final product is NAD+-bound red semiquinone of the enzyme.  相似文献   

14.
A novel approach is described for generating reactive oxidizing centers in heme proteins, with zinc hemoglobin (Zn Hb) and zinc cytochrome c (Zn cyt c) used as examples. The reaction of 3Zn* Hb with [CoIII(NH3)5 Cl]2+, and of 3Zn* cyt c with methyl viologen are described. In the case of Zn Hb the cation radical produced decays with a rate constant of k3 = 2400s-1. Using this value the rate of the reaction (formula; see text) can be calculated to be 4500s-1.  相似文献   

15.
Various oxidized mono/di/tri/poly saccharides were studied as potential hemoglobin (Hb) cross-linkers in order to produce oxygen carriers with high oxygen affinities (low P(50)'s) and high molecular weights (therefore lower macromolecular diffusivities compared to tetrameric Hb). Such physical properties were desired to produce polymerized hemoglobins (PolyHbs) with oxygen release profiles similar to that of human blood, as was demonstrated in work by Winslow (1). In this present study, bovine hemoglobin was cross-linked with a variety of oxidized (ring-opened) saccharides, which resulted in cross-linked Hb species ranging in size from 64 to 6400 kDa (depending on the particular oxidized saccharide used in the reaction) and P(50)'s ranging from 6 to 15 mmHg. A parallel synthetic approach was used to synthesize these carbohydrate-hemoglobin conjugates, and asymmetric flow field-flow fractionation (AFFF) coupled with multi-angle static light scattering (MASLS) was used to measure the absolute molecular weight distribution of these PolyHb dispersions. Cross-linking reactions were conducted at two pHs (6 and 8), with larger cross-linked Hb species produced at pH 8 (where hydrolysis was most likely to occur between glycosidic bonds linking adjacent saccharide rings) rather than at pH 6. The largest molecular weight species formed from these reactions consisted of Hb cross-linked with ring-opened lactose, maltose, methylglucopyranoside, sucrose, trehalose, and 15 kDa and 71 kDa dextran at high pH (pH 8). The most promising Hb cross-linker was methylglucopyranoside, which resulted in very large cross-linked Hb species, with low P(50)'s and lower methemoglobin (metHb) levels compared to the other Hb cross-linking reagents.  相似文献   

16.
The kinetics of the reversible binding of cyanide by the ferric cytochrome c' from Chromatium vinosum have been studied over the pH range 6.9-9.6. The reaction is extremely slow at neutral pH compared to the reactions of other high-spin ferric heme proteins with cyanide. The observed bimolecular rate constant at pH 7.0 is 2.25 X 10(-3) M-1 s-1, which is approximately 10(7)-fold slower than that for peroxidases, approximately 10(5)-fold slower than those for hemoglobin and myoglobin, and approximately 10(2)-fold to approximately 10(3)-fold slower than that recently reported for the Glycera dibranchiata hemoglobin, which has anomalously slow cyanide rate constants of 4.91 X 10(-1), 3.02 X 10(-1), and 1.82 M-1 s-1 for components II, III, and IV, respectively [Mintorovitch, J., & Satterlee, J. D. (1988) Biochemistry 27, 8045-8050; Mintorovitch, J., Van Pelt, D., & Satterlee, J. D. (1989) Biochemistry 28, 6099-6104]. The unusual ligand binding property of this cytochrome c' is proposed to be associated with a severely hindered heme coordination site. Cyanide binding is also characterized by a nonlinear cyanide concentration dependence of the observed rate constant at higher pH values, which is interpreted as involving a change in the rate-determining step associated with the formation of an intermediate complex between the cytochrome c' and cyanide prior to coordination. The pH dependence of both the binding constant for the formation of the intermediate complex and the association rate constant for the subsequent coordination to the heme can be attributed to the ionization of HCN, where cyanide ion binding is the predominant process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
When O2 was injected into an anaerobic suspension of valinomycin-treated rat liver mitochondria inhibited with rotenone, antimycin, and myxothiazol, a small amount of O2 (0.23-0.33 ng-atom of O/mg of protein) was reduced extremely rapidly (within the 2 s time-resolution of the oxygen electrode). The subsequent steady-state rate of flow of electrons to oxygen was very low [less than 3 nequiv. X s-1 X (g of mitochondrial protein)-1]. In the presence of valinomycin there was a rapid ejection of protons synchronous with the rapid phase of O2 consumption corresponding to 0.38-0.61 nequiv. of H+ X (mg of mitochondrial protein)-1. When valinomycin was replaced by carbonyl cyanide p-trifluoromethoxyphenylhydrazone (FCCP) there was a rapid alkalification of the medium corresponding to 0.20-0.42 nequiv. of H+ X (mg of mitochondrial protein)-1. When 2 mM-Fe(CN)6(4-) was present to re-reduce endogenous cytochrome c, O2 consumption was still biphasic but the second phase of O2 consumption was very much more rapid [600 nequiv. X s-1 X (g of protein)-1], and resulted in the virtually complete consumption of the O2 in the pulse within 4 s. With 60 microM-Ru(NH3)6(2+) as reductant, O2 consumption was even faster [1200 nequiv. X s-1 X (g of protein)-1]. In a medium containing 150 mM-choline chloride with Ru(NH3)6(2+) as reductant, the proton per reducing equivalent stoichiometry (delta H+O/e-) was +0.95 in the presence of valinomycin and -0.94 in the presence of FCCP. In choline chloride medium containing Ru(NH3)6(2+) and valinomycin, there was an uptake of K+ ions corresponding to 1.86 K+/e-. It is concluded that nearly 1 proton is translocated outwards through cytochrome oxidase per oxidizing equivalent injected in this medium. In low ionic strength sucrose-based medium, with Ru(NH3)6(2+) as reductant, delta H+O/e- was 1.05 in the presence of valinomycin, and -0.71 in the presence of FCCP. It is concluded that the translocation of protons is accompanied by net acid production in this medium.  相似文献   

18.
In reaction centers from Rhodopseudomonas viridis that contain a single quinone, the decay of the photo-oxidized primary donor, P+, was found to be biphasic when the bound, donor cytochromes were chemically oxidized by ferricyanide. The ratio of the two phases was dependent on pH with an apparent pK of 7.6. A fast phase, which dominated at high pH (t1/2 = 1 ms at pH 9.5), corresponded to the expected charge recombination of P+ and the primary acceptor QA-. A much slower phase dominated at low pH and was shown to arise from a slow reduction of P+ by ferrocyanide in reaction centers where QA- has been rapidly oxidized by ferricyanide. The rate of QA- oxidation was linear with respect to ferricyanide activity and was strongly pH-dependent. The second-order rate constant, corrected for the activity coefficient of ferricyanide, approached a maximum of 2 X 10(8) M-1 X s-1 at low pH, but decreased steadily as the pH was raised above a pK of 5.8, indicating that a protonated state of the reaction center was involved. The slow reduction of P+ by ferrocyanide was also second-order, with a maximum rate constant at low pH of 8 X 10(5) M-1 X s-1 corrected for the activity coefficient of ferrocyanide. This rate also decreased at higher pH, with a pK of 7.4, indicating that ferrocyanide also was most reactive with a protonated form of the reaction center. The oxidation of QA- by ferricyanide was unaffected by the presence of o-phenanthroline, implying that access to QA- was not via the QB-binding site. In reaction centers supplemented with ubiquinone, oxidation of reduced secondary quinone, QB-, by ferricyanide was observed but was substantially slower than that for QA-. It is suggested that Q-B may be oxidized via QA so that the rate is modulated by the equilibrium constant for QA-QB in equilibrium with QAQB-.  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.
The oxidation of uric acid and 3-N-ribosyluric acid by hydrogen peroxide and methemoglobin was stimulated by the addition of sodium nitrite, which alone has no effect on the urates. The urates were not oxidized by either hydrogen peroxide alone or hydrogen peroxide and sodium nitrite unless methemoglobin was present. t-Butyl hydroperoxide also oxidized the urates in the presence of methemoglobin, but the reaction was not stimulated by sodium nitrite. The addition of either sodium azide or potassium cyanide reduced the rate of the reaction with either hydrogen peroxide or t-butyl hydroperoxide both in the presence and absence of sodium nitrite. Possible explanations for the stimulation by nitrite of peroxide-induced degradation of urates are presented.  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