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.
2.
《BBA》1986,848(3):402-410
Effects of temperature and dehydration on the efficiency of electron transfer from membrane-bound high-potential cytochromes ch to the reaction-center bacteriochlorophyll (P-890) in Ectothiorhodospira shaposhnikovii have been studied. A kinetic analysis of the cytochrome oxidation suggests that there are at least two conformational states of the ch-P-890 complex, of which only one allows photoinduced electron transfer from cytochrome to P-890+. Lowering the temperature of dehydration leads to a change in the proportion of the populations in the two conformations. The observed 2-fold deceleration of cytochrome oxidation can be related only to the diminution of the amount of photoactive cytochromes per reaction center. The rate constant for the transfer of an electron from cytochrome ch to bacteriochlorophyll is 2.8 · 105 s−1 and is independent of temperature and dehydration (as estimated within the accuracy of the experiments). The effects produced by low temperature and dehydration are completely reversible. The thermodynamic parameters of the transition of the cytochrome from the nontransfer to electron-transfer conformation were estimated. For room temperature (+ 20°C) in chromatophore preparations, ΔG = −5.4 kJ · M−1, ΔH = 60 kJ · M−1, ΔS = 0.22 kJ · M−1 · K−1. For Triton X-100 subchromatophore preparations, the absolute values of the above parameters are significantly lower: ΔG = −2.8 kJ · M−1, ΔH = 18 kJ · M−1, and ΔS = 0.075 kJ · M−1 · K−1. To a larger extent, the above parameters are diminished for chromatophore preparations in an 80% glycerol solution: ΔG = −1.7 kJ · M−1, ΔH = 6 kJ · M−1, ΔS = 0.025 kJ · M−1 · K−1. The data suggest the hydrophobic character of the forces that maintain the P-890-ch complex in the electron-transfer conformation. The results obtained suggest that electron tunneling within the complex cannot occur until a specific conformational configuration of the complex is formed. The efficiency of cytochrome ch oxidation is determined by the temperature, the degree of dehydration and the environmental conditions, whereas the transfer of an electron itself in the electron-transfer configuration is essentially independent of temperature and hydration.  相似文献   

3.
《FEBS letters》1985,182(1):31-33
Three types of proteoliposome containing mitochondrial H+-ATPase have been prepared: Mg2+-‘free’, one-side and two-side Mg2+-containing proteoliposomes. The ATPase activity as well as its sensitivity to oligomycin or N,N'-dicyclohexylcarbodiimide of the three proteoliposome preparations has been compared. They decreased in the order : L ·(H+-ATPase)+Mg2+ > L · (H+-ATPase)+Mg2+ > L · (H+-ATPase)−Mg2+. The fluidity of the proteoliposomes has also been compared by fluorescence polarization probes diphenylhexatriene (DPH) or 7-(9-anthroyloxy)stearic acid (7-AS). The degree of polarization for DPH in these proteoliposomes decreased in the order: L · (H+-ATPase)+Mg2+ > L · (H+-ATPase)+Mg2+ > L · (H+-ATPase)−Mg2+, while that for 7-AS: L · (H+-ATPase)+Mg2+ ≈ L · (H+-ATPase)+Mg2+ > L · (H+-ATPase)−Mg2+.Lipid fluidityMitochondrial H+-ATPaseOne-side Mg2+ effectTwo-side Mg2+ effectLipid-protein interaction  相似文献   

4.
《BBA》2006,1757(9-10):1133-1143
In cytochrome c oxidase, oxido-reductions of heme a/CuA and heme a3/CuB are cooperatively linked to proton transfer at acid/base groups in the enzyme. H+/e cooperative linkage at Fea3/CuB is envisaged to be involved in proton pump mechanisms confined to the binuclear center. Models have also been proposed which involve a role in proton pumping of cooperative H+/e linkage at heme a (and CuA). Observations will be presented on: (i) proton consumption in the reduction of molecular oxygen to H2O in soluble bovine heart cytochrome c oxidase; (ii) proton release/uptake associated with anaerobic oxidation/reduction of heme a/CuA and heme a3/CuB in the soluble oxidase; (iii) H+ release in the external phase (i.e. H+ pumping) associated with the oxidative (R  O transition), reductive (O  R transition) and a full catalytic cycle (R  O  R transition) of membrane-reconstituted cytochrome c oxidase. A model is presented in which cooperative H+/e linkage at heme a/CuA and heme a3/CuB with acid/base clusters, C1 and C2 respectively, and protonmotive steps of the reduction of O2 to water are involved in proton pumping.  相似文献   

5.
The stoichiometry of hydroxylation reactions catalyzed by cytochrome P-450 was studied in a reconstituted enzyme system containing the highly purified cytochrome from phenobarbital-induced rabbit liver microsomes. Hydrogen peroxide was shown to be formed in the reconstituted system in the presence of NADPH and oxygen; the amount of peroxide produced varied with the substrated added. NADPH oxidation, oxygen consumption, and total product formation (sum of hydroxylated compound and hydrogen peroxide) were shown to be equimolar when cyclohexane, benzphetamine, or dimethylaniline served as the substrate. The stoichiometry observed represents the sum of two activities associated with cytochrome P-450. These are (1) hydroxylase activity: NADPH + H+ + O2 + RH → NADP+ + H2O + ROH; and (2) oxidase activity: NADPH + H+ + O2 → NADP+ + H2O2. Benzylamphetamine (desmethylbenzphetamine) acts as a pseudosubstrate in that it stimulates peroxide formation to the same extent as the parent compound (benzphetamine), but does not undergo hydroxylation. Accordingly, when benzylamphetamine alone is added in control experiments to correct for the NADPH and O2 consumption not associated with benzphetamine hydroxylation, the expected 1:1:1 stoichiometry for NADPH oxidation, O2 consumption, and formaldehyde formation in the hydroxylation reaction is observed.  相似文献   

6.
The reactions of NO2 with both oxidized and reduced cytochrome c at pH 7.2 and 7.4, respectively, and with N-acetyltyrosine amide and N-acetyltryptophan amide at pH 7.3 were studied by pulse radiolysis at 23 °C. NO2 oxidizes N-acetyltyrosine amide and N-acetyltryptophan amide with rate constants of (3.1±0.3)×105 and (1.1±0.1)×106 M−1 s−1, respectively. With iron(III)cytochrome c, the reaction involves only its amino acids, because no changes in the visible spectrum of cytochrome c are observed. The second-order rate constant is (5.8±0.7)×106 M−1 s−1 at pH 7.2. NO2 oxidizes iron(II)cytochrome c with a second-order rate constant of (6.6±0.5)×107 M−1 s−1 at pH 7.4; formation of iron(III)cytochrome c is quantitative. Based on these rate constants, we propose that the reaction with iron(II)cytochrome c proceeds via a mechanism in which 90% of NO2 oxidizes the iron center directly—most probably via reaction at the solvent-accessible heme edge—whereas 10% oxidizes the amino acid residues to the corresponding radicals, which, in turn, oxidize iron(II). Iron(II)cytochrome c is also oxidized by peroxynitrite in the presence of CO2 to iron(III)cytochrome c, with a yield of ~60% relative to peroxynitrite. Our results indicate that, in vivo, NO2 will attack preferentially the reduced form of cytochrome c; protein damage is expected to be marginal, the consequence of formation of amino acid radicals on iron(III)cytochrome c.  相似文献   

7.
《BBA》1985,806(2):320-330
Two membrane-associated cytochromes, cytochrome cm-553 and cytochrome cm-552, were derived from Nitrosomonas europaea. The major c-type cytochrome, cytochrome cm-553, accounted for 92% of the c heme found in the membrane. It had absorption maxima at 410 nm in the oxidized form and at 417, 523 and 553 nm in the dithionite reduced form. Cytochrome cm-552 possessed absorption maxima at 409 nm in the oxidized form, at 421, 522 and 552 in the dithionite reduced form, and at 418 in the dithionite reduced plus CO form. The concentration and cellular distribution of the two c-type membrane cytochromes, hydroxylamine oxidoreductase and cytochromes c-552, c-554, and a were determined. Over 95% of the soluble cytochromes (hydroxylamine oxidoreductase cytochromes and c-552 and c-554) were periplasmic, whereas cytochrome cm-553, cytochrome cm-552 and cytochrome a were associated with the cell membrane. The outer membrane and cytoplasm were devoid of cytochromes. The extracytoplasmic location of the proton-yielding hydroxylamine oxidizing system (NH2OH ™ HNO + 2H+ + 2e) may contribute to an energy-linked proton gradient. The heme concentrations of hydroxylamine oxidoreductase and cytochromes c-552, c-554, cm-553, cm-552 and a were approx. 2.4, 1.2, 0.3, 1.3, 0.1 and 1.1 nmol/mg cell protein, respectively. The corresponding molar ratios of heme were 22:11:2.9:12:1.0:10. The enzyme or cytochrome concentrations for hydroxylamine oxidoreductase and cytochromes c-552, c-554, cm-553, cm-552 and a were approx. 0.13, 1.05, 0.09, 0.63, 0.055 and 0.56 nmol/mg cell protein, respectively. The corresponding molar ratios were 0.24:2.2:0.16:1.2:0.1:1.0.  相似文献   

8.
To test the possibility of inorganic carbon limitation of the marine unicellular alga Emiliania huxleyi (Lohmann) Hay and Mohler, its carbon acquisition was measured as a function of the different chemical species of inorganic carbon present in the medium. Because these different species are interdependent and covary in any experiment in which the speciation is changed, a set of experiments was performed to produce a multidimensional carbon uptake scheme for photosynthesis and calcification. This scheme shows that CO2 that is used for photosynthesis comes from two sources. The CO2 in seawater supports a modest rate of photosynthesis. The HCO is the major substrate for photosynthesis by intracellular production of CO2 (HCO+ H+→ CO2+ H2O → CH2O + O2). This use of HCO is possible because of the simultaneous calcification using a second HCO, which provides the required proton (HCO+ Ca2+→ CaCO3+ H+). The HCO is the only substrate for calcification. By distinguishing the two sources of CO2 used in photosynthesis, it was shown that E. huxleyi has a K½ for external CO2 of “only” 1.9 ± 0.5 μM (and a Vmax of 2.4 ± 0.1 pmol·cell−1·d−1). Thus, in seawater that is in equilibrium with the atmosphere ([CO2]= 14 μM, [HCO]= 1920 μM, at fCO2= 360 μatm, pH = 8, T = 15° C), photosynthesis is 90% saturated with external CO2. Under the same conditions, the rate of photosynthesis is doubled by the calcification route of CO2 supply (from 2.1 to 4.5 pmol·cell−1·d−1). However, photosynthesis is not fully saturated, as calcification has a K½ for HCO of 3256 ± 1402 μM and a Vmax of 6.4 ± 1.8 pmol·cell−1·d−1. The H+ that is produced during calcification is used with an efficiency of 0.97 ± 0.08, leading to the conclusion that it is used intracellularly. A maximum efficiency of 0.88 can be expected, as NO uptake generates a H+ sink (OH source) for the cell. The success of E. huxleyi as a coccolithophorid may be related to the efficient coupling between H+ generation in calcification and CO2 fixation in photosynthesis.  相似文献   

9.
The title complex undergoes decomposition in acidic aqueous solution resulting in equimolar concentration of aquapentaamminecobalt(III) and hexa- aquacobalt(II). The kinetic studies over the ranges of 0.048 M ⩽ [H+] ⩽ 0.385 M, 25 ⩽ θc ⩽ 41.5°C and at I = 0.5 M reveals that the intricate mechanism involves protonation equilibrium of the title complex, followed by a rate determining bridge cleavage. The further follow-up reaction is a fast electron transfer process to form products. The rate expression derived from the mechanism is kobs = k1K1[H+]/(1 + K1[H+]) where the values of k, and K, are found to be 8.9 × 10−4 s−1 and 3.5 M−1 respectively at 25 °C. The results are compared with that obtained for the decomposition reactions of mononuclear aquaammine complexes of cobalt(III).  相似文献   

10.
The oxidation-reduction reaction of horse heart cytochrome c and cytochrome c (552, Thermus thermophilus), which is highly thermoresistant, was studied by temperature-jump method. Ferrohexacyanide was used as reductant.
Thermodynamic and activation parameters of the reaction obtained for both cytochromes were compared with each other. The results of this showed that (1) the redox potential of cytochrome c-552,+0.19 V, is markedly less than that of horse heart cytochrome c. (2) ?Hox3 of cytochrome c-552 is considerably lower than that of horse heart cytochrome c. (3) ?Hox3 and ?Sred3 of cytoochrome c-552 are more negative than those of horse heart cytochrome c. (4) kred of cytochrome c-552 is much lower than that of horse heart cytochrome c at room temperature.  相似文献   

11.
Peter R. Rich  Peter Heathcote 《BBA》1983,723(2):332-340
(i) Purified bovine heart mitochondrial cytochrome b-c1 complex (ubiquinone-cytochrome c oxidoreductase) and photosynthetic reaction centres isolated from Rhodopseudomonas sphaeroides strain R-26 have been incorporated into lipid vesicles. In the presence of cytochrome c and ubiquinone-2, light activation caused a cyclic electron transfer involving both components. (2) Since cytochrome c is added outside the vesicles, it is both reduced by the cytochrome b-c1 complex and oxidised by the reaction centre on the outside of the vesicles. Ubiquinone-2, however, is reduced by the reaction centres at a site in contact with the inside of the vesicles, but the reduced form, ubiquinol-2, is oxidised by the cytochrome b-c1 complex at a site in contact with the outer aqueous phase. (3) In the presence of valinomycin plus K+, initiation of cyclic electron flow causes protons to move from inside the vesicles to the outer medium and the H+2e? ratio was calculated to be close to 4.  相似文献   

12.
The kinetics of electron transfer between the isolated enzymes of cytochrome c1 and cytochrome c have been investigated using the stopped-flow technique. The reaction between ferrocytochrome c1 and ferricytochrome c is fast; the second-order rate constant (k1) is 3.0 · 107 M?1 · s?1 at low ionic strength (I = 223 mM, 10°C). The value of this rate constant decreases to 1.8 · 105 M?1 · s?1 upon increasing the ionic strength to 1.13 M. The ionic strength dependence of the electron transfer between cytochrome c1 and cytochrome c implies the involvement of electrostatic interactions in the reaction between both cytochromes. In addition to a general influence of ionic strength, specific anion effects are found for phosphate, chloride and morpholinosulphonate. These anions appear to inhibit the reaction between cytochrome c1 and cytochrome c by binding of these anions to the cytochrome c molecule. Such a phenomenon is not observed for cacodylate. At an ionic strength of 1.02 M, the second-order rate constants for the reaction between ferrocytochrome c1 and ferricytochrome c and the reverse reaction are k1 = 2.4 · 105 M?1 · s?1 and k?1 = 3.3 · 105 M?1 · s?1, respectively (450 mM potassium phosphate, pH 7.0, 1% Tween 20, 10°C). The ‘equilibrium’ constant calculated from the rate constants (0.73) is equal to the constant determined from equilibrium studies. Moreover, it is shown that at this ionic strength, the concentrations of intermediary complexes are very low and that the value of the equilibrium constant is independent of ionic strength. These data can be fitted into the following simple reaction scheme: cytochrome c2+1 + cytochrome c3+ai cytochrome c3+1 + cytochrome c2+.  相似文献   

13.
We previously reported that high micromolar concentrations of nitric oxide were able to oxidize mitochondrial cytochrome c at physiological pH, producing nitroxyl anion (Sharpe and Cooper, 1998 Biochem. J. 332, 9–19). However, the subsequent re-evaluation of the redox potential of the NO/NO- couple suggests that this reaction is thermodynamically unfavored. We now show that the oxidation is oxygen-concentration dependent and non stoichiometric. We conclude that the effect is due to an oxidant species produced during the aerobic decay of nitric oxide to nitrite and nitrate. The species is most probably nitrogen dioxide, NO2? a well-known biologically active oxidant. A simple kinetic model of NO autoxidation is able to explain the extent of cytochrome c oxidation assuming a rate constant of 3 × 106 M-1 s-1 for the reaction of NO2? with ferrocytochrome c. The importance of NO2? was confirmed by the addition of scavengers such as urate and ferrocyanide. These convert NO2? into products (urate radical and ferricyanide) that rapidly oxidize cytochrome c and hence greatly enhance the extent of oxidation observed. The present study does not support the previous hypothesis that NO and cytochrome c can generate appreciable amounts of nitroxyl ions (NO- or HNO) or of peroxynitrite.  相似文献   

14.
《BBA》2020,1861(1):148087
Electron bifurcating, [FeFe]-hydrogenases are recently described members of the hydrogenase family and catalyze a combination of exergonic and endergonic electron exchanges between three carriers (2 ferredoxinred + NAD(P)H + 3 H+ = 2 ferredoxinox + NAD(P)+ + 2 H2). A thermodynamic analysis of the bifurcating, [FeFe]-hydrogenase reaction, using electron path-independent variables, quantified potential biological roles of the reaction without requiring enzyme details. The bifurcating [FeFe]-hydrogenase reaction, like all bifurcating reactions, can be written as a sum of two non-bifurcating reactions. Therefore, the thermodynamic properties of the bifurcating reaction can never exceed the properties of the individual, non-bifurcating, reactions. The bifurcating [FeFe]-hydrogenase reaction has three competitive properties: 1) enabling NAD(P)H-driven proton reduction at pH2 higher than the concurrent operation of the two, non-bifurcating reactions, 2) oxidation of NAD(P)H and ferredoxin simultaneously in a 1:1 ratio, both are produced during typical glucose fermentations, and 3) enhanced energy conservation (~10 kJ mol−1 H2) relative to concurrent operation of the two, non-bifurcating reactions. Our analysis demonstrated ferredoxin E°′ largely determines the sensitivity of the bifurcating reaction to pH2, modulation of the reduced/oxidized electron carrier ratios contributed less to equilibria shifts. Hydrogenase thermodynamics data were integrated with typical and non-typical glycolysis pathways to evaluate achieving the ‘Thauer limit’ (4 H2 per glucose) as a function of temperature and pH2. For instance, the bifurcating [FeFe]-hydrogenase reaction permits the Thauer limit at 60 °C if pH 2 ≤ ~10 mbar. The results also predict Archaea, expressing a non-typical glycolysis pathway, would not benefit from a bifurcating [FeFe]-hydrogenase reaction; interestingly, no Archaea have been observed experimentally with a [FeFe]-hydrogenase enzyme.  相似文献   

15.
16.
Sulfur sources capable of replacing sulfide were surveyed for biomethanation from H2 and CO2 by thermoautotrophic methanogen, Methanobacterium thermoautotrophicum. Among sulfur containing compounds tested, l-cysteine, thiosulfate and coenzyme M gave poor growth when added as sulfur sources, whereas simultaneous addition of two sulfur sources, l-cysteine+thiosulfate, l-cysteine+l-methionine or l-cysteine+coenzyme M stimulated the growth.In a pressure-controlled fermentor system developed to obtain stoichiometry between input and output gases, the ratio of H2 and CO2 consumption to CH4 production was almost stoichiometric, and when l-cysteine and thiosulfate or l-methionine were used in place of sulfide (control) similar growth patterns were observed. In a culture with continuous supply of substrates gases (1.3 vvm) and sulfur sources of 1 mM l-cysteine+2 mM thiosulfate, specific growth rate and specific methane production rate were 0.35 h and 3.24 l g−1h−1, respectively, compared to 0.22 h−1 and 5.76 l gh−1 with Na2 S.  相似文献   

17.
Oligochitosan has been proved to trigger plant cell death. To gain some insights into the mechanisms of oligochitosan-induced cell death, the nature of oligochitosan-induced cell death and the role of calcium (Ca2+), nitric oxide (NO) and hydrogen peroxide (H2O2) were studied in tobacco suspension cells. Oligochitosan-induced cell death occurred in cytoplasmic shrinkage, phosphatidylserine externalization, chromatin condensation, TUNEL-positive nuclei, cytochrome c release and induction of programmed cell death (PCD)-related gene hsr203J, suggesting the activation of PCD pathway. Pretreatment cells with cyclosporin A, resulted in reducing oligochitosan-induced cytochrome c release and cell death, indicating oligochitosan-induced PCD was mediated by cytochrome c. In the early stage, cells undergoing PCD showed an immediate burst in free cytosolic Ca2+ ([Ca2+]cyt) elevation, NO and H2O2 production. Further study showed that these three signals were involved in oligochitosan-induced PCD, while Ca2+ and NO played a negative role in this process by modulating cytochrome c release.  相似文献   

18.
Petronijevic T., Rogers W. P. and Sommerville R. I. 1985. Carbonic acid as the host signal for the development of parasitic stages of nematodes. International Journal for Parasitology15: 661–667. This paper gives results on which may be based an identification of the component of the system CO2 + H2O ai H2CO3 ai H+ HCO3? which acts as the stimulus from the animal host for some nematodes. Using infective juveniles of Nematospiroides dubius and Haemonchus contortus, the effects on exsheathment of (1) low pCO2 values, (2) the presence of carbonic anhydrase in the stimulating medium, and (3) the inhibition of carbonic anhydrase within the juveniles have been examined. The results lead to the suggestion that it is the “readily available” undissociated H2CO3, or H2CO3 + HCO3? which is the critical factor in the stimulus for development. The wide range of [H+]s over which “readily available” H2CO3 is present in physiological environments suggests that this host signal may be important for infection with many species.  相似文献   

19.
Schumaker KS  Sze H 《Plant physiology》1985,79(4):1111-1117
Two types of ATP-dependent calcium (Ca2+) transport systems were detected in sealed microsomal vesicles from oat roots. Approximately 80% of the total Ca2+ uptake was associated with vesicles of 1.11 grams per cubic centimeter and was insensitive to vanadate or azide, but inhibited by NO3. The remaining 20% was vanadate-sensitive and mostly associated with the endoplasmic reticulum, as the transport activity comigrated with an endoplasmic reticulum marker (antimycin A-insensitive NADH cytochrome c reductase), which was shifted from 1.11 to 1.20 grams per cubic centimeter by Mg2+.

Like the tonoplast H+-ATPase activity, vanadate-insensitive Ca2+ accumulation was stimulated by 20 millimolar Cl and inhibited by 10 micromolar 4,4′-diisothiocyano-2,2′-stilbene disulfonic acid or 50 micromolar N,N′-dicyclohexylcarbodiimide. This Ca2+ transport system had an apparent Km for Mg-ATP of 0.24 millimolar similar to the tonoplast ATPase. The vanadate-insensitive Ca2+ transport was abolished by compounds that eliminated a pH gradient and Ca2+ dissipated a pH gradient (acid inside) generated by the tonoplast-type H+-ATPase. These results provide compelling evidence that a pH gradient generated by the H+-ATPase drives Ca2+ accumulation into right-side-out tonoplast vesicles via a Ca2+/H+ antiport. This transport system was saturable with respect to Ca2+ (Km apparent = 14 micromolar). The Ca2+/H+ antiport operated independently of the H+-ATPase since an artifically imposed pH gradient (acid inside) could also drive Ca2+ accumulation. Ca2+ transport by this system may be one major way in which vacuoles function in Ca2+ homeostasis in the cytoplasm of plant cells.

  相似文献   

20.
The kinetics and mechanism of a linear trihydroxamic acid siderophore (deferriferrioxamine B, H4DFB+) ligand exchange with Al(H2O)63+ to form mono(deferriferrioxamine B)aluminum(III) (Al(H2O)4H3DFB)3+ have been investigated at 25 °C over the [H+] range 0.001−1.0 M and I = 2.0 M (HClO4/NaClO4) by 27Al NMR. Kinetic results are consistent with Al(H2O)4(H3DFB)3+ formation and dissociation proceeding through a parallel path mechanistic scheme involving Al(H2O)63+(k2/k−1) and Al(H2O)5(OH)2+(k2/k−2) where k1 = 0.13 M−1 s−1, k−1 = 8.7 × 10−3 M−1 s−1, k2 = 2.7 × 103 M−1 s−1, and k−2 = 9.6 × 10−4 s−1. Relative complex formation rates at Al(H2O)63+ and Al(H2O)5OH2+, and comparison with kinetic data for a series of synthetic hydroxamic acids, suggest that an interchange mechanism is operative. These results are also discussed in relation to kinetic data for the corresponding iron(III)-deferriferrioxamine B system.  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