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 46 毫秒
1.
Chloroplasts which were rapidly isolated from illuminated leaves showed activity of ATP hydrolysis at a level much higher than that of the dark control. Under the high-intensity illumination or under repetitive flash excitation, the activated chloroplasts synthesized more ATP than those with a low ATP hydrolysis activity. Δ\?gmH+ formed under repetitive flashes was smaller in the activated chloroplasts than in the inactive chloroplasts. The inhibition of ATP yield per flash by valinomycin or nigericin in the presence of K+ was stronger in the inactive chloroplasts than in the activated chloroplast. ATP synthesis in the activated chloroplasts seems to have a lower Δ\?gmH+ threshold.  相似文献   

2.
ADP and Pi-loaded membrane vesicles from l-malate-grown Bacillus alcalophilus synthesized ATP upon energization with ascorbateN,N,N′,N′-tetramethyl-p-phenylenediamine. ATP synthesis occurred over a range of external pH from 6.0 to 11.0, under conditions in which the total protonmotive force Δ\?gmH+ was as low as ?30 mV. The phosphate potentials (ΔGp) were calculated to be 11 and 12 kcal/mol at pH 10.5 and 9.0, respectively, whereas the Δ\?gmH+ values in vesicles at these two pH values were quite different (?40 ± 20 mV at pH 10.5 and ?125 ± 20 mV at pH 9.0). ATP synthesis was inhibited by KCN, gramicidin, and by N,N′-dicyclohexylcarbodiimide. Inward translocation of protons, concomitant with ATP synthesis, was demonstrated using direct pH monitoring and fluorescence methods. No dependence upon the presence of Na+ or K+ was found. Thus, ATP synthesis in B. alcalophilus appears to involve a proton-translocating ATPase which functions at low Δ\?gmH+.  相似文献   

3.
G.F. Azzone  T. Pozzan  E. Viola  P. Arslan 《BBA》1978,501(2):317-329
1. The aerobic uptake of inorganic ions, such as 86Rb+ or 125I?, by submitochondrial particles, is about one order of magnitude lower than the uptake of organic ions, such as acridines or 8-anilino-1-naphthalene sulphonate. The values of ΔpH, the transmembrane pH differential, and Δψ, the transmembrane membrane potential are between 60 and 100 mV when calculated on the inorganic ions and between 150 and 240 mV when calculated on the organic ions. The discrepancy between the ΔpH and Δψ values from organic and inorganic ions is large at high but not at low ion/protein ratios.2. In the absence of weak bases and strong acids the values of Δ\?gmH, the proton electrochemical potential difference, are close to 100 mV and the magnitude of ΔpH and Δψ are similar. Weak bases decrease ΔpH and enhance Δψ. Strong acids decrease Δψ and enhance ΔpH. Interchangeability of ΔpH with Δψ occurs at low concentrations of weak bases and strong acids. High concentrations of weak bases and strong acids cause depression of Δ\?gmH.3. Concentrations of weak bases capable of abolishing ΔpH, do not affect ATP synthesis. Concentrations of strong acids capable of abolishing Δψ affect only slightly ATP synthesis. Concentrations of weak bases and strong acids capable of causing a decline of ΔpH + Δψ inhibit ATP synthesis.4. Depression of Δ\?gmH is paralleled by inhibition of ATP synthesis and decline of ΔGp, the phosphate potential. Abolition of ATP synthesis occurs only when Δ\?gmH is below 20 mV. The ΔGp\?gmH ratio increases hyperbolically with the decrease of Δ\?gmH.  相似文献   

4.
Luit Slooten  Adriaan Nuyten 《BBA》1981,638(2):305-312
(1) The ATPase enzyme in untreated chromatophores from Rhodospirillum rubrum is in a low-activity state (designated as E°). It can be activated by application of a transmembrane Δ\?gmH+ generated by light-induced electron transport, or by application of acid-base jumps. (2) After rapid dissipation of the light-induced Δ\?gmH+, the active state of the ATPase enzyme decays (in the absence of added substrates or products) to a low-activity state (designated as E′), with a half-time of the order of 2–4 min. This state differs from E° in that E′ (but not E°) can be rapidly reactivated by addition of substrate, but only when the Mg2+ concentration is kept below 20–30 μM. Since this is characteristic of an activated enzyme containing tightly bound ADP (Slooten, L. and Nuyten, A. (1981) Biochim. Biophys. Acta 638, 313–326), it is suggested that release of endogenous, tightly bound ADP is one of the factors involved in activation of the ATPase enzyme.  相似文献   

5.
The Km(app) of ADP for photophosphorylation in lettuce chloroplasts was measured both at various light intensities and in the presence of various uncoupler (nigericin + K+) concentrations. Lowering the light intensity results in both, a decrease in the rate of phosphorylation and a several fold decrease in the Km(app) of ADP for the reaction. However, when increasing concentrations of the uncoupler nigericin + K+ are employed, the rate of photophosphorylation is decreased but a several-fold increase in the Km(app) of ADP for the reaction is observed. The results are discussed in terms of the chemiosmotic hypothesis. It is suggested that these effects might indicate the existence of a mechanism controlling the rate of ATP formation which is different than the formation of the electrochemical gradient.  相似文献   

6.
(1) Treatment of (Na+ + K+)-ATPase from rabbit kidney outer medulla with the γ-35S labeled thio-analogue of ATP in the presence of Na+ + Mg2+ and the absence of K+ leads to thiophosphorylation of the enzyme. The Km value for [γ-S]ATP is 2.2 μM and for Na+ 4.2 mM at 22°C. Thiophosphorylation is a sigmoidal function of the Na+ concentration, yielding a Hill coefficient nH = 2.6. (2) The thio-analogue (Km = 35 μM) can also support overall (Na+ + K+)-ATPase activity, but Vmax at 37°C is only 1.3 γmol · (mg protein)? · h?1 or 0.09% of the specific activity for ATP (Km = 0.43 mM). (3) The thiophosphoenzyme intermediate, like the natural phosphoenzyme, is sensitive to hydroxylamine, indicating that it also is an acylphosphate. However, the thiophosphoenzyme, unlike the phosphoenzyme, is acid labile at temperatures as low as 0°C. The acid-denatured thiophosphoenzyme has optimal stability at pH 5–6. (4) The thiophosphorylation capacity of the enzyme is equal to its phosphorylation capacity, indicating the same number of sites. Phosphorylation by ATP excludes thiophosphorylation, suggesting that the two substrates compete for the same phosphorylation site. (5) The (apparent) rate constants of thiophosphorylation (0.4 s?1 vs. 180 s?1), spontaneous dethiophosphorylation (0.04 s?1 vs. 0.5 s?1) and K+-stimulated dethiophosphorylation (0.54 s?1 vs. 230 s?1) are much lower than those for the corresponding reactions based on ATP. (6) In contrast to the phosphoenzyme, the thiophosphoenzyme is ADP-sensitive (with an apparent rate constant in ADP-induced dethiophosphorylation of 0.35 s?1, KmADP = 48 μM at 0.1 mM ATP) and is relatively K+-insensitve. The Km for K+ in dethiophosphorylation is 0.9 mM and in dephosphorylation 0.09 mM. The thiophosphoenzyme appears to be for 75–90% in the ADP-sensitive E1-conformation.  相似文献   

7.
The stoichiometry of free NADPH oxidation in phenobarbital induced rabbit liver microsomes was measured by means of registering the rates of NADPH, H+ and O2 consumption and O2? and H2O2 production. ΔO2?:ΔH2O2 ratio is approximately I indicating that about half H2O2 results from O2? dismutation, the second half being formed directly. ΔNADPH:ΔH2O2 and ΔO2:ΔH2O2 ratios exceed I and therefore another product of the reaction is water. The fact that the ratio (ΔNADPH-ΔH2O2):(ΔO2-ΔH2O2) is 2 allows one to consider direct 4-electron O2 reduction as the major way of water formation rather than endogenous substrate hydroxylation.  相似文献   

8.
Cytoplasmic membrane vesicles isolated from Escherichia coli take up dansyl-galactoside, a fluorescent competitive inhibitor of lactose transport, to much lower levels than lactose. An initial interpretation, based on the study of the fluorescent changes accompanying the energy-dependent uptake, was that it represented a one-to-one specific binding to the lac carrier protein which was not followed by transport. Recently, on the basis of a new estimation of the number of lac carrier in the membrane, it has been advanced that the uptake of dansyl-galactoside represents a nonspecific binding on the inner surface of the membrane following transport. We discriminate between the two interpretations by comparing the effects of lactose and dansyl-galactoside uptake on the electrochemical gradient of protons (Δ\?gmH+), generated by the oxidation of substrates, and on the uptake of proline. Indeed, it is known that the rate of lactose transport is such that it leads, as a consequence of the lactose/H+ symport, to an observable decrease of Δ\?gmH+, and secondary to this decrease to an inhibition of the uptake of proline transported at much lower rate. We show that the rates of uptake of lactose and dansyl-galactoside by the membrane vesicles are similar; yet the uptake of dansyl-galactoside does not lead to the uncoupling effects which are associated with the uptake of lactose. We discuss the possible reasons for the absence of this uncoupling effect, and we conclude that our data are incompatible with the notion that the energy-dependent uptake of dansyl-galactoside is associated with an active transport involving a dansyl-galactoside/H+ symport. On the contrary, the data substantiate the initial interpretation that the energy-dependent uptake of dansyl-galactoside reflects the binding to the lac carrier not followed by transport.  相似文献   

9.
Extant photosynthetic organisms all appear to use transmembrane H+ fluxes as the coupling agent in the use of light energy in ATP synthesis. In the steady-state there is a large H+ free energy difference across the coupling membrane, and when this is reflected as a light-induced change in pH of the phase (cytosol or stroma) containing the enzymes of carbon assimilation, the H+ transport can have an informational role in activating and inactivating enzymes.The earliest organisms probably lived fermentatively (substrate-level phosphorylation) in an anaerobic environment provided with organic solutes synthesised abiotically. There are good reasons for believing that one of the earliest primary active transport systems (interconverting chemical and electrical/osmotic energy) was an H+ extrusion pump powered by ATP or PPi. Its initial function was extrusion of excess H+ from the fermenting cells, and the support of a number of co-transport processes. The earliest energetic use of light energy is envisaged as being the energization of an alternative H+ extrusion pump, with bacteriorhodopsin or (bacterio-) chlorophyll as the pigment. The former type of cyclic photoredox system (Halobacterium-type) is simpler than the latter: a “pre-respiratory” chemical redox H+ pump may have preceded the (bacterio-) chlorophyll-based process. Any of these H+ pumps could spare the use of fermentative ATP in powering active H+ efflux and would thus have been favoured as fermentative substrates became scarce; eventually the larger ΔμH+ generated by the light-powered H+ pump was used to drive the ATP-powered H+ pump backwards and thus generate ATP with light as the ultimate energy source.Scarcity of suitable reductants for biosynthesis as life proliferated provided a selective impetus for a non-cyclic photoredox system which could use light energy to generate a low-potential reductant at the expense of more readily available higher-potential reductants. The non-cyclic photoredox system is not possible in its simplest form (with all the redox energy coming from excitation energy of one or more photoreactions) in the bacteriorhodopsin line of evolution. Such a simple photoredox system is found in the Chlorobiaceae; even if (as seems likely) the non-cyclic photoredox process generates a ΔμH+ (and thus, potentially, ATP), some of the ATP needed for CO2 fixation and cell growth must be generated by a cyclic photoredox system.In the extant purple bacteria the generation of low-potential reductant involves a non-cyclic photoredox pathway which produces a reductant unable to reduce NAD+; the “energy gap” is spanned by “reverse electron transfer” which uses energy from a ΔμH+. It is not clear if this energetic requirement for the H+ gradient can be quantitatively satisfied from a non-cyclic photoredox H+ transport; it is certain that there is a major requirement for cyclic photoredox H+ pumping in these organisms.The photosynthetic bacteria are today restricted to reducing (low Eh) environments similar to those found in the early, anoxic earth; they are unable to use very weak reductants as donors for non-cyclic photoredox processes. As the sources of even weakly reducing donors (other than H2O) on the primitive earth were depleted the two photoreactions scheme of extant O2-producers evolved by modification of the bacterial photoreaction. This non-cyclic photoredox process is definitely H+-translocating and the role of cyclic photoredox processes in ATP generation in O2-evolvers is smaller than in photosynthetic bacteria.In parallel with the biochemical and biophysical changes in the photosystems there was a morphological evolution, with an increasing tendency for “internalisation” of the photoredox processes (originally present in the plasma membrane, as in extant Chlorobineae) into thylakoids (as in most Rhodospirillineae, Cyanobacteria and in all eukaryotes). With a plasmalemma-located photoredox system, and the constraints of a fixed, alkaline external pH and the cytoplasmic pH of 7–8, the ΔμH+ would be generated largely as an electrical P.D. The presence of a phase (intrathylakoid space) with a “negotiable pH” would permit the generation and use of a ΔμH+ largely present as a pH gradient.In both cases illumination can cause an increase in cytoplasmic (stromal) pH over the dark value; this is an important aspect of the regulation of “phototrophic” and “heterotrophic” enzyme systems in the light and in the dark. However, it is argued that these differences in pH are not absolutely light-dependent unless they depend upon some more uniquely light-dependent signal, probably based on a redox component only generated in the light.  相似文献   

10.
(1) H+/electron acceptor ratios have been determined with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing endogenous substrates during reduction of O2, NO?2 or N2O. Under optimal H+-translocation conditions, the ratios H+O, H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were 6.0–6.3, 4.02, 5.79 and 3.37, respectively. (2) With ascorbate/N,N,N′,N′-tetramethyl-p-phenylenediamine as exogenous substrate, addition of NO?2 or N2O to an anaerobic cell suspension resulted in rapid alkalinization of the outer bulk medium. H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were ?0.84, ?2.33 and ?1.90, respectively. (3) The H+oxidant ratios, mentioned in item 2, were not altered in the presence of valinomycinK+ and the triphenylmethylphosphonium cation. (4) A simplified scheme of electron transport to O2, NO?2 and N2O is presented which shows a periplasmic orientation of the nitrite reductase as well as the nitrous oxide reductase. Electrons destined for NO?2, N2O or O2 pass two H+-translocating sites. The H+electron acceptor ratios predicted by this scheme are in good agreement with the experimental values.  相似文献   

11.
A capacitor microphone was used to measure the enthalpy and volume changes that accompany the electron transfer reactions, PQAhv P+Q?A and PQAQBhv P+QAQ?B, following flash excitation of photosynthetic reaction centers isolated from Rhodopseudomonas sphaeroides. P is a bacteriochlorophyll dimer (P-870), and QA and QB are ubiquinones. In reaction centers containing only QA, the enthalpy of P+Q?A is very close to that of the PQA ground state (ΔHr = 0.05 ± 0.03 eV). The free energy of about 0.65 eV that is captured in the photochemical reaction evidently takes the form of a substantial entropy decrease. In contrast, the formation of P+QAQ?B in reaction centers containing both quinones has a ΔHr of 0.32 ± 0.02 eV. The entropy change must be near zero in this case. In the presence of o-phenanthroline, which blocks electron transfer between Q?A and QB, ΔHr for forming P+Q?AQB is 0.13 ± 0.03 eV. The influence of flash-induced proton uptake on the results was investigated, and the ΔHr values given above were measured under conditions that minimized this influence. Although the reductions of QA and QB involve very different changes in enthalpy and entropy, both reactions are accompanied by a similar volume decrease of about 20 ml/mol. The contraction probably reflects electrostriction caused by the charges on P+ and Q?A or Q?B.  相似文献   

12.
In an accompanying publication by Duckwitz-Peterlein, Eilenberger and Overath ((1977) Biochim. Biophys. Acta 469, 311–325) it is shown that the exchange of lipid molecules between negatively charged vesicles consisting of total phospholipid extracts from Escherichia coli occurs by the transfer of single lipid monomers or small micelles through the water. Here a kinetic interpretation is presented in terms of a rate constant, k?, for the escape of lipid molecules from the vesicle bilayer into the water. The evaluated rate constants are k?P = (0.86 ± 0.05) · 10?5s?1 and k?E = (1.09 ± 0.13) · 10?6s?1 for phospholipid molecules with trans-Δ9-hexadecenoate and trans-Δ9-octadecenoate, respectively, as the predominant acyl chain component. The rate constants are discussed in terms of the acyl chain and polar head group composition of the lipids.  相似文献   

13.
14.
The oxygen dependence of cellular energy metabolism.   总被引:14,自引:0,他引:14  
Suspensions of cultured C 1300 neuroblastoma cells, sarcoma 180 ascites tumor cells, and Tetrahymena pyriformis cells were used to study the oxygen dependence of cellular energy metabolism. Cellular respiration was found to be almost independent of oxygen tension to values of less than 20 μm with an apparent Km for oxygen of less than 1 μm. In contrast, the reduction of mitochondrial cytochrome c was found to be dependent on oxygen tension at all values from 240 μm downward. Oxygen dependence was also observed in terms of cellular energy metabolism expressed as adenosine triphosphate and adenosine diphosphate concentrations. These data provide direct evidence that in intact cells mitochondrial oxidative phosphorylation is oxygen dependent throughout the physiological range of oxygen tension (air saturation and below). The respiratory rate is maintained constant when the oxygen tension is lowered by decreasing values of the cytosolic [ATP][ADP][Pi] and intramitochondrial [NAD]+][NADH] because these regulatory parameters adjust to maintain a constant rate of ATP synthesis. The lack of oxygen dependence in the respiratory rate means that the rate of cellular ATP utilization is essentially oxygen independent until the mitochondria can no longer synthesize ATP at the required rate and [ATP][ADP][Pi].  相似文献   

15.
The permeability of the lysosomal membrane to small anions and cations was studied at 37°C and pH 7.0 in a lysosomal-mitochondrial fraction isolated from the liver of untreated rats. The extent of osmotic lysis following ion influx was used as a measure of ion permeancy. In order to preserve electroneutrality, anion influx was coupled to an influx of K+ in the presence of valinomycin, and cation influx was coupled to an efflux of H+ using the protonophore 3-tert-butyl-5,2′-dichloro-4′-nitrosalicilylanilide. Lysosomal lysis was monitored by observing the loss of latency of two lysosomal hydrolases.The order of permeability of the lysosomal membrane to anions was found to be SCN? > I? > CH3COO? > Cl? ≈ HCO?3 ≈ Pi > SO42? and that to cations Cs+ > K+ > Na+ > H+. These orders are largely in agreement with the lyotropic series of anions and cations.The implications of these findings for the mechanism by means of which a low intralysosomal pH is produced and maintained are discussed.  相似文献   

16.
A thermodynamic characterization of the Na+-H+ exchange system in Halobacterium halobium was carried out by evaluating the relevant phenomenological parameters derived from potential-jump measurements. The experiments were performed with sub-bacterial particles devoid of the purple membrane, in 1 M NaCl, 2 M KCl, and at pH 6.5–7.0. Jumps in either pH or pNa were brought about in the external medium, at zero electric potential difference across the membrane, and the resulting relaxation kinetics of protons and sodium flows were measured. It was found that the relaxation kinetics of the proton flow caused by a pH-jump follow a single exponential decay, and that the relaxation kinetics of both the proton and the sodium flows caused by a pNa-jump also follow single exponential decay patterns. In addition, it was found that the decay constants for the proton flow caused by a pH-jump and a pNa-jump have the same numerical value. The physical meaning of the decay constants has been elucidated in terms of the phenomenological coefficients (mobilities) and the buffering capacities of the system. The phenomenological coefficients for the Na+-H+ flows were determined as differential quantities. The value obtained for the total proton permeability through the particle membrane via all available channels, LH = (?JH +pH)Δψ,ΔpNa, was in the range of 850–1150 nmol H+·(mg protein)?1·h?1·(pH unit)?1 for four different preparations; for the total Na+ permeability, LNa = (?JNa+pNa)Δψ,ΔpH, it was 1620–2500 nmol Na+·(mg protein)?1·h?1·(pNa unit)?1; and for the proton ‘cross-permeability’, LHNa = (?JH+pNa)Δψ,ΔpH, it was 220–580 nmol H+·(mg protein)?1·h?1·(pNa unit)?1, for different preparations. From the above phenomenological parameters, the following quantities have been calculated: the degree of coupling (q), the maximal efficiency of Na+-H+ exchange (ηmax), the flow and force efficacies (?) of the above exchange, and the admissible range for the values of the molecular stoichiometry parameter (r). We found q ? 0.4; ηmax ? 5%; 0.36 ? r ? 2; ?JNa+ ? 1.3 · 105μmol · (RT unit)?1 at JNa = 1 μmolNa+ · (mgprotein)?1 · h?1; and ?ΔpNa ? 5 · 104 ΔpNa · (mg protein) · h · (RT unit)?1 at ΔpNa = 1 unit, for different preparations.  相似文献   

17.
R.L. Pan  S. Izawa 《BBA》1979,547(2):311-319
NH2OH-treated, non-water-splitting chloroplasts can oxidize H2O2 to O2 through Photosystem II at substantial rates (100–250 μequiv · h?1 · mg?1 chlorophyll with 5 mM H2O2) using 2,5-dimethyl-p-benzoquinone as an electron acceptor in the presence of the plastoquinone antagonist dibromothymoquinone. This H2O2 → Photosystem II → dimethylquinone reaction supports phosphorylation with a Pe2 ratio of 0.25–0.35 and proton uptake with H+e values of 0.67 (pH 8)–0.85 (pH 6). These are close to the Pe2 value of 0.3–0.38 and the H+e values of 0.7–0.93 found in parallel experiments for the H2O → Photosystem II → dimethylquinone reaction in untreated chloroplasts. Semi-quantitative data are also presented which show that the donor → Photosystem II → dibromothymoquinone (→O2) reaction can support phosphorylation when the donor used is a proton-releasing reductant (benzidine, catechol) but not when it is a non-proton carrier (I?, ferrocyanide).  相似文献   

18.
19.
(1) The total phospholipid content of a gradient purified (K+ + H+)-ATPase preparation from pig gastric mucosa is 105 μmol per 100 mg protein, and consists of 29% sphingomyelin, 29% phosphatidylcholine, 28% phosphatidylethanolamine, 10% phosphatidylserine and 4% phosphatidylinositol. The cholesterol content corresponds to 50 μmol per 100 mg protein. (2) Treatment with phospholipase C (from Clostridium welchii and Bacillus cereus) results in an immediate decrease of the phosphate content. Up to 50% of the phospholipids are hydrolyzed by each phospholipase C preparation alone, without further hydrolysis by increased phospholipase concentration or prolonged incubation time. Combined treatment with the two phospholipase C preparations, sequentially or simultaneously, hydrolyzes up to 65% of the phospholipids. (3) The (K+ + H+)-ATPase and K+ stimulated p-nitrophenylphosphatase activities are decreased proportionally with the total phospholipid content, indicating that these enzyme activities are dependent on phospholipids. (4) Phospholipase C treatment does not change optimal pH, Km value for ATP and temperature dependence of the gastric (K+ + H+)-ATPase, but slightly decreases the Ka value for K+. (5) Phospholipase C treatment lowers the AdoPP[NH]P binding and phosphorylation capacities, suggesting that inactivation occurs primarily on the substrate binding level. (6) Most of the results can be understood by assuming that hydrolysis of the phospholipids by phospholipase C leads to aggregation of the membrane protein molecules and complete inactivation of the aggregated ATPase molecules.  相似文献   

20.
The cell-free preparations from autotrophieally grown Pseudomonas saccharophila catalyzed the process of electron transport from H2 or various other organic electron donors to either O2 or NO3? with concomitant ATP generation. The respective PO ratios with H2 and NADH were 0.63 and 0.73, the respective PNO3? ratios were 0.57 and 0.54. In contrast, the PO and PNO3? ratios with succinate were 0.18 and 0.11, respectively. ATP formation coupled to the oxidation of ascorbate, in the absence or presence of added N,N,N′,N′-tetramethyl-p-phenylenediamine or cytochrome c, could not be detected. Various uncouplers inhibited phosphorylation with either O2 or NO3? as terminal electron acceptors without affecting the oxidation of H2 or other substrates. The NADH oxidation at the expense of O2 or NO3? reduction as well as the associated phosphorylation were inhibited by rotenone and amytal. The aerobic and anaerobic H2 oxidation and coupled ATP synthesis, on the other hand, was unaffected by the flavoprotein inhibitors as well as by the NADH trapping system. The NADH, H2, and succinate-linked electron transport to O2 or NO3? and the associated phosphorylations were sensitive, however, to antimycin A or 2-n-nonyl-4-hydroxyquino-line-N-oxide, and cyanide or azide. The data indicated that although the phosphorylation sites 1 and II were associated with NADH oxidation by O2 or NO3?, the energy conservation coupled to H2 oxidation under aerobic or anaerobic conditions appeared to involve site II only.  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