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.
Pyridoxamine (pyridoxine) 5′-phosphate oxidase (EC 1.4.3.5) purified from rabbit liver is competitively inhibited by the reaction product, pyridoxal 5′-phosphate. The Ki, 3 μM, is considerably lower than the Km for either natural substrate (18 and 24 μM for pyridoxamine 5′-phosphate and 25 and 16 μM for pyridoxine 5′-phosphate in 0.2 M potassium phosphate at pH 8 and 7, respectively). The Ki determined using a 10% rabbit liver homogenate is the same as that for the pure enzyme; hence, product inhibition invivo is probably not diminished significantly by other cellular components. Similar determinations for a 10% rat liver homogenate also show strong inhibition by pyridoxal 5′-phosphate. Since the reported liver content of free or loosely bound pyridoxal 5′-phosphate is greater than Ki, the oxidase in liver is probably associated with pyridoxal 5′-phosphate. These results also suggest that product inhibition of pyridoxamine-P oxidase may regulate the invivo rate of pyridoxal 5′-phosphate formation.  相似文献   

2.
In Acanthamoeba castellanii mitochondria, the apparent affinity values of alternative oxidase for oxygen were much lower than those for cytochrome c oxidase. For unstimulated alternative oxidase, the KMox values were around 4-5 μM both in mitochondria oxidizing 1 mM external NADH or 10 mM succinate. For alternative oxidase fully stimulated by 1 mM GMP, the KKMox values were markedly different when compared to those in the absence of GMP and they varied when different respiratory substrates were oxidized (KMox was around 1.2 μM for succinate and around 11 μM for NADH). Thus, with succinate as a reducing substrate, the activation of alternative oxidase (with GMP) resulted in the oxidation of the ubiquinone pool, and a corresponding decrease in KMox. However, when external NADH was oxidized, the ubiquinone pool was further reduced (albeit slightly) with alternative oxidase activation, and the KMox increased dramatically. Thus, the apparent affinity of alternative oxidase for oxygen decreased when the ubiquinone reduction level increased either by changing the activator or the respiratory substrate availability.  相似文献   

3.
The single glutathione S-transferase (EC 2.5.1.18) present in rat erythrocytes was purified to apparent homogeneity by affinity chromatography on glutathione-Sepharose and hydroxyapatite chromatography. Approx. 1.86 mg enzyme is found in 100 ml packed erythrocytes and accounts for about 0.01% of total soluble protein. The native enzyme (Mr 48 000) displays a pI of 5.9 and appears to possess a homodimeric structure with a subunit of Mr 23 500. Enzyme activities with ethacrynic acid and cumene hydroperoxide were 24 and 3%, respectively, of that with 1-chloro-2,4-dinitrobenzene. The Km values for 1-chloro-2,4-dinitrobenzene and glutathione were 1.0 and 0.142 mM, respectively. The concentrations of certain compounds required to produce 50% inhibition (I50) were as follows: 12 μM bromosulphophthalein, 34 μM S-hexylglutathione, 339 μM oxidized glutathione and 1.5 mM cholate. Bromosulphophthalein was a noncompetitive inhibitor with respect to 1-chloro-2,4-dinitrobenzene (Ki = 8 μM) and glutathione (Kis = 4 μM; Kii = 11.5 μM) while S-hexylglutathione was competitive with glutathione (Ki = 5 μM).  相似文献   

4.
Neither salicylate nor ibuprofen was a substrate or inhibitor of the long-chain fatty acid: CoA ligase. In contrast, all three xenobiotic-metabolizing medium-chain fatty acid:CoA ligases (XL-I, XL-II, and XL-III) had activity toward salicylate. The Km value for salicylate was similar for all three forms (2 to 3 μM), but XL-II and XL-III had higher activity at Vmax. For ibuprofen, only XL-III catalyzed its activation, and it had a Km for ibuprofen of 36 μM. Studies of salicylate inhibition of XL-I, XL-II, and XL-III revealed that it inhibited the benzoate activity of all three forms with K1 values of ca. 2 μM, which is in agreement with the Km values obtained with salicylate as substrate. Kinetic analysis revealed that salicylate conjugation by all three forms is characterized by substrate inhibition when salicylate exceeds ca. 20 μM. Substrate inhibition was more extensive with XL-I and XL-III. Previous work on the ligases employed assay concentrations of salicylate in the range of 0.1 to 1.0 mM, which are clearly inhibitory, particularly toward XL-I and XL-III. Thus, activity was not properly measured in previous studies, which accounts for the fact that salicylate conjugation was only found with one form, which is most likely XL-II since it has the highest Vmax activity and shows the least amount of substrate inhibition. Studies with ibuprofen indicated that it inhibited XL-I, XL-II, and XL-III, with K1 values being in the range of 75–125 μM. The short-chain ligase was inhibited by both salicylate and ibuprofen with K1 values of 93 and 84 μM, respectively. It was concluded that pharmacological doses of salicylate, but not ibuprofen, will affect the metabolism of medium-chain fatty acids and carboxylic acid xenobiotics and that the previously described mitochondrial ibuprofen:CoA ligase activity is attributable to XL-III. © 1996 John Wiley & Sons, Inc.  相似文献   

5.
1. Cell-free extracts of the marine bacterium Beneckea natriegens, derived by sonication, were separated into particulate and supernatant fractions by centrifugation at 150 000 × g.2. NADH, succinate, d(?)- and l(+)-lactate oxidase and dehydrogenase activities were located in the particles, with 2- to 3-fold increases in specific activity over the cell free extract. The d(?)- and l(+)-lactate dehydrogenases were NAD+ and NADP+ independent. Ascorbate-N,N,N′,N′-tetramethylphenylenediamine (TMPD) oxidase was also present in the particulate fraction; it was 7–12 times more active than the physiological substrate oxidases.3. Ascorbate-TMPD oxidase was completely inhibited by 10 μM cyanide. Succinate, NADH, d(?)-lactate and l(+)-lactate oxidases were inhibited in a biphasic manner, with 10 μM cyanide causing only 10–50 % inhibition; further inhibition required more than 0.5 mM cyanide, and 10 mM cyanide caused over 90 % inhibition. Low sulphide (5 μM) and azide (2 mM) concentrations also totally inhibited ascorbate-TMPD oxidase, but only partially inhibited the other oxidases. High concentrations of sulphide but not azide caused a second phase inhibition of NADH, succinate, d(?)-lactate and l(+)-lactate oxidases.4. Low oxidase activities of the physiological substrates, obtained by using non-saturating substrate concentrations, were more inhibited by 10 μM cyanide and 2 mM azide than high oxidase rates, yet ascorbate-TMPD oxidase was completely inhibited by 10 μM cyanide over a wide range of rates of oxidation.5. These results indicate terminal branching of the respiratory system. Ascorbate-TMPD is oxidised by one pathway only, whilst NADH, succinate, d(?)-lactate and l(+)-lactate are oxidised via both pathways. Respiration of the latter substrates occurs preferentially by the pathway associated with ascorbate-TMPD oxidase and which is sensitive to low concentrations of cyanide, azide and sulphide.6. The apparent Km for O2 for each of the two pathways was detected using ascorbate-TMPD and NADH or succinate plus 10 μM cyanide respectively. The former pathway had an apparent Km of 8–17 (average 10.6) μM and the latter 2.2–4.0 (average 3.0) μM O2.  相似文献   

6.
The enzymatic production of α-dehydrobiotin (α-DHB), an antibiotic, from biotinyl-CoA using acyl-CoA oxidase and from biotin using a coupling system of biotinyl-CoA synthetase and acyl-CoA oxidase was developed. Acyl-CoA oxidase was found to show activity for biotinyl-CoA. Km and Vmax values of acyl-CoA oxidase for biotinyl-CoA were 75 μM and 3.92 μmol min−1 mg−1, respectively. Optimum reaction conditions for the α-DHB production from biotin were examined. The maximum production of α-DHB (4.29 μmol ml−1) was obtained, when the reaction was carried out at 30°C for 36 h in a mixture consisting of 100 mM potassium phosphate buffer (pH 8.0), 20 mM biotin, 20 mM ATP, 60 mM CoA, 20 mM MgCl2, 2 units of biotinyl-CoA synthetase, 90 units of acyl-CoA oxidase and 25 units of catalase in a total volume of 0.6 ml under aerobic conditions. The product was purified from 14 ml of the reaction mixture and 10 mg of crystals with white needle form were obtained. From NMR, mass spectra and other physical analyses, this compound was identified as (+)-trans-α-DHB.  相似文献   

7.
(1) Investigation of the relationship between the detergent concentration and steady-state and pre-steady-state kinetics of cytochrome c oxidase proved to be a valid approach in the study of protein-detergent interaction. (2) Laurylmaltoside, sodium cholate and Triton X-100 influenced the kinetics of cytochrome c oxidase cooperatively at detergent concentrations near their critical micelle concentration. This mode of interaction reflects disaggregation of the oxidase as a result of cooperative binding of the detergent. (3) Addition of increasing concentrations of Tween-80 to the aggregated enzyme caused a more gradual decrease in aggregation of the oxidase, which did not result in a change in activity of the enzyme. This suggests that aggregation of cytochrome c oxidase occurs in a highly regular manner in which no catalytic sites are shielded off. (4) Oxidase aggregates present at detergent concentrations below the critical micelle concentration of laurylmaltoside and Triton X-100 showed considerable activity. Their kinetics were equal to those of the oxidase in Tween-80, suggesting that the protein molecules are aligned in a similar way in all oligomers. Aggregates present in low concentrations of sodium cholate showed turnover rates that were twice as low as those observed with other aggregates. (5) Solubilisation of the oxidase by sodium cholate or Triton X-100 resulted in almost complete inhibition of enzymic activity, whereas the association rate of ferrocytochrome c was almost equal to that found for monomeric oxidase in laurylmaltoside. These results are in agreement with a mixed-type inhibition.  相似文献   

8.
9.
Choline oxidase catalyzes the oxidation of choline to glycine-betaine, with betaine-aldehyde as intermediate and molecular oxygen as primary electron acceptor. This study reports on the inhibitory effects of triarylmethanes (cationic malachite green; neutral leukomalachite green), phenoxazines (cationic, meldola blue and nile blue; neutral nile red) and a structurally-related phenothiazine (methylene blue) on choline oxidase, assayed at 25°C in 50 mM MOPS buffer, pH 7, using choline as substrate. Methylene B acted as a competitive inhibitor with Ki = 74 ± 7.2 μM, pointing to the choline–binding site of the enzyme as a target site. Nile B caused noncompetitive inhibition of enzyme activity with Ki = 20 ± 4.5 μM. In contrast to methylene B and nile B, malachite G and meldola B caused complex, nonlinear inhibition of choline oxidase, with estimated Ki values in the micromolar range. The difference in kinetic pattern was ascribed to the differential ability of the dyes to interact (and interfere) with the flavin cofactor, generating different perturbations in the steady-state balance of the catalytic process.  相似文献   

10.
  • 1.1. The effects of prostaglandin (PG) E1, and I2 analogs (OP-41483 and OP-2507) on the Superoxide generation of human neutrophil NADPH oxidase (EC 1.6.99.6) in both whole-cell and cell-free systems were investigated.
  • 2.2. In a whole-cell system, OP-2507 inhibited the Superoxide generation by neutrophils exposed to phorbol myristate acetate concentration-dependently through its superoxide-scavenging action.
  • 3.3. The concentration of the drug required for 50% inhibition of the oxidase (IC50) was 21 μM.
  • 4.4. In a cell-free system, however, the drug in concentrations of < 100 μM did not inhibit the activation of NADPH oxidase by sodium dodecyl sulfate because of its inactivation by the detergent.
  • 5.5. Although PGE1 and OP-41483 did not inhibit the Superoxide production by stimulated neutrophils in a whole-cell system, they both inhibited the activation of NADPH oxidase in a cell-free system concentration-dependently, with IC50 values of 44 and 170 μM, respectively.
  • 6.6. In addition, in the cell-free system, the Km value for NADPH of the oxidase was unchanged by PGE1.
  • 7.7. The results suggest that the PGI2 analog, OP-2507, is a possible superoxide-scavenger and that PGE1 inhibits the NADPH oxidase activation by sodium dodecyl sulfate in a cell-free system concentration-dependently.
  相似文献   

11.
A modified multiple binding equilibria treatment is presented that allows determination of thermodynamic parameters of the interaction of phospholipids with integral membrane proteins solubilized in excess detergent. Lipid binding is modeled as a series of exchange reactions between lipid molecules and detergent molecules at the hydrophobic protein surface. A general equation is derived which expresses a relative association constant (K) and the total number of contact sites at the lipid-protein interface (N) in terms of experimentally measurable variables. A useful simplification of the general equation occurs when the amount of detergent is high relative to the total number of lipid binding sites in the sample. Computer simulations show that in cases we have examined there appears to be an experimentally accessible range of detergent to protein molar ratios where the approximation at high detergent is useful for analyzing experimental data. This model is used to examine the competition between cholate and spin-labeled phospholipids for the hydrophobic surfaces of bovine heart cytochrome c oxidase. We find, for example, that K = 12 +/- 2 for phosphatidylcholine relative to cholate (i.e., the cholate molecules are relatively easily displaced by membrane lipids). This helps to explain the experimental observation that cholate is an effective detergent both for solubilizing cytochrome c oxidase and for reconstituting this protein into a defined lipid bilayer environment. An excess of cholate readily displaces almost all of the native phospholipids, and the protein is dispersed in cholate micelles. However, when phospholipids are added back, the cholate molecules at the protein surface are replaced because of the higher relative binding of the phospholipids. Observed differences between the behavior of phosphatidylcholine and phosphatidylglycerol suggest that reconstitution in cholate is a selective process in which detergent molecules in localized areas on the protein surface are more readily displaced by certain phospholipids.  相似文献   

12.
Fibrinogen showed essentially no binding (KD>1 mM ) to platelet αIIbβ3 integrin in solution in the presence of Triton or octylglucoside above critical micellar concentrations. Under these conditions the integrin was an αβ monomer. After removal of the detergent from the Triton containing buffer (25 mM Tris/HCl;, 150 mM NaCl, 1 mM CaCl2, 1 mM MgCl2, pH 7.4) the integrin formed aggregates with hexamers as the most prominent species, as demonstrated by analytical ultracentrifugation and electron microscopy. Tracer sedimentation equilibrium experiments indicate that fibrinogen binds to the integrin aggregates, but with a surprisingly large KD (at least 3 μM ). This value is 10- to 100-fold higher than values determined by solid phase assays or with integrins reconstituted onto lipid bilayers.  相似文献   

13.
Cultured human lymphoblastoid cells take up taurine from the medium by two processes: 1) a temperature-dependent, Na+-dependent, saturable “active”-transport system and 2) diffusion. The active transport has properties similar to those reported for taurine transport by other tissues. Apparent Km is about 25 μM and Vmax about 7.2 pmol/min/106 cells; saturation occurs at 100 μM taurine. Uptake is competitively inhibited by the β-amino acids hypotaurine (50% inhibition at 44 μM) and β-alanine (50% at 152 μM), as measured at 50 μM taurine. Taurocyamine inhibits 50% at 260 μM. Chlorpromazine and imipramine are strong uncompetitive inhibitors, giving 50% inhibition at 26 μM and 115 μM, respectively; at these concentrations cellular viability per se is not affected. Ouabain inhibits 40–50% over a concentration range of 4–500 μM. Diffusion of taurine into the cells is proportional to concentration up to 20 mM. However, at the concentration of taurine in human plasma, 40–100 μM, active transport would provide 90% of the taurine taken up.  相似文献   

14.
Guanylate cyclase activity is present in both soluble and particulate fractions of homogenates of mouse cerebellum and retina. Soluble guanylate cyclases in cerebellum and retina have an apparent Km for GTP of approx 40 and 70 μM, respectively; are stimulated by Ca2+ and Mg2+ in the presence of low Mn2+; and do not respond to NaN3, NH2OH or detergent. The particulate guanylate cyclase found in brain has an apparent Km GTP of 237 7mu;M, is not stimulated by Ca2+ or Mg2+ in the presence of low Mn2+, but is stimulated by NaN3, NH2OH, and detergent. In particulate fractions of normal retina, guanylate cyclase has two apparent Km GTP values (42 and 225 μM); has higher activity at low concentrations of Mn2+ (0.5 mM) than at high concentrations (5.0 mM); is inhibited by Ca2+; and does not respond to NaN3, NH2OH, or detergent. Retinas essentially devoid of photoreceptor cells (from mice with photoreceptor dystrophy) have soluble guanylate cyclase activity which is similar to that in normal retina, but have only 4% as much particulate guanylate cyclase activity. This residual particulate guanylate cyclase has an apparent Km GTP value of 392 μM and other properties similar to particulate guanylate cyclase from brain. These data indicate the presence of three distinguishable guanylate cyclases in CNS: (1) a soluble enzyme present in both brain and retina: (2) a particulate enzyme which is also present in brain and in the inner or neural retina: and (3) another particulate enzyme which is apparently unique and confined to retinal photoreceptor cells.  相似文献   

15.
The properties of NADH-dependent Fe3+-EDTA reductase in plasma membranes (PM) from roots of iron-deficient and -sufficient tomato plants [Lycopersicon esculentum L. (Mill.) cv. Abunda] were examined. Iron deficiency resulted in a 3-fold increase of in vivo root iron-chelate reductase activity with a Km (Fe3+-EDTA) of 230 μM. In purified root PM, average specific activities of ferric chelate reductase of 410 and 254 nmol Fe (mg protein)?1 min?1 were obtained for iron-deficient and -sufficient plants, respectively. In both cases, the PM-bound activity showed a pH optimum at pH 6.8. Activity depended on NADH and not on NADPH and on the presence of detergent. The activity was inhibited 40-50% by superoxide dismutase (EC 1.15.1.1) and ca 30% by oxygen. Kinetic analysis of the membrane-bound enzyme revealed a Km (Fe3+-EDTA) of ca 200 μM for both iron-stressed and -sufficient plants. For NADH, Km values around 230 μM were obtained. The ferric chelate reductase could be solubilised from salt-washed PM with Triton X-100 at a protein:detergent ratio of 1:2.8 (w/w). The Triton-soluble fraction revealed one enzyme-stained band in native polyacrylamide electrophoresis. Although the membranes showed no nitrate reductase (NR; EC 1.6.6.1) activity, anti-spinach NR immunoglobulin G (IgG) recognized a 54 kDa band both in the PM and the Triton-soluble fraction, but not in the enzymatically active material obtained from the native gel. No evidence could be found for the synthesis of a new, biochemically distinct PM-bound ferric chelate reductase under iron deficiency, which might be identified as the so-called Turbo reductase. It is concluded that iron deficiency in tomato induces increased expression of a ferric chelate reductase in root PM, which is already present in iron-sufficient plants and probably also in plants, which do not contain the Turbo reductase, like the grasses. The iron reductase is not identical with the recently reported PM-associated nitrate reductase.  相似文献   

16.
The effect of temperature on the silicon limited growth and nutrient kinetics of Stephanodiscus minutus Grun. was examined using batch and semicontinuous culture methods. Short-term batch culture methods gave maximum growth rates which were essentially constant over the temperature range of 10° to 20°C (μ3= 0.71–0.80 d?1). The half-saturation constant for growth (Ks) was significantly lowest at 10°C (Ks= 0.31 μM Si; 0.22–0.41), and higher at both 15°C (Ks= 1.03 μM Si; 0.68–1.47) and 20°C (Ks= 0.88 μM Si; 0.60–1.22). Two methods were used to evaluate the semicontinuous experiments. The Droop relationship showed that the minimum cell quota was about 1.50 × 10?7 nmol Si cell?1, but there was much overlap in the results at all three temperatures. The Monod growth relationship for the semicontinuous experiments gave estimates of Ks which were lowest at 15°C (Ks= 0.12 μM Si), and higher at 10°C (Ks= 0.68 μM Si) and 20°C (Ks= 1.24 μM Si), although 95% confidence intervals overlapped. The maximum growth rate estimates for the semicontinuous experiments were similar at 10° and 15°, and higher at 20°C, but the number of points used in making the calculations makes the results less reliable than those from batch cultures. Generally, there were no consistent significant differences in the silicon limited growth of S. minutus over the temperature range studied. Our values of Ks for S. minutus are the lowest recorded for a freshwater diatom, and are consistent with the distribution of this species in nature. Generally, this species becomes abundant in areas with high phosphorus loading and very low silicon levels (low Si:P loading rates). Stephanodiscus species are also fossil indicators of eutrophication in north temperate lakes.  相似文献   

17.
D. discoideum contains kinetically distinguishable cell surface cAMP binding sites. One class, S, is slowly dissociating and has high affinity for cAMP (Kd = 15 nM, t12 = 15 s). A second class is fast dissociating (t12 about 1 s) and is composed of high affinity binding sites H (Kd ≈ 60 nM), and low affinity binding sites L (Kd = ≈ 450 nM) which interconvert during the binding reaction. Guanine nucleotides affect these three binding types in membranes prepared by shearing D.discoideum cells through Nucleopore filters. The affinity of S for cAMP is reduced by guanine nucleotides from 13 nM to 25 nM, and the number of S-sites is reduced about 50%. The number of fast dissociating sites is not altered by guanine nucleotides, but these sites are mainly in the low affinity state. Half-maximal effects are obtained at about 1 μM GTP, 2 μM GDP and 10 μM Gpp(NH)p(guanyl-5′-yl-imidodiphosphate); ATP and ADP are without effect up to 1 mM. These results indicate that D.discoideum cells have a functionally active guanine nucleotide binding protein involved in the transduction of extracellular cAMP signals via cell surface cAMP receptors.  相似文献   

18.
1. The intermediate structures formed during dialysis of mixtures of cholate, phospholipid and cytochrome c oxidase were analysed by gel chromatography and electron microscopy. Measurements of trapped phosphate and the degree of respiratory control were used to assess the integrity of the vesicular structures formed. Protein orientation in the bilayer was monitored by the accessibility of cytochrome c to cytochrome c oxidase. 2. The results indicate that proteoliposome formation by the detergent-dialysis procedure takes place in three distinct stages. In the first stage, cholate/phospholipid and cholate/phospholipid/protein micelles coexist in solution and grow in size as the detergent is slowly removed. At a detergent/phospholipid molar ratio of about 0.2, micelle fusion results in the formation of large bilayer aggregates permeable to both phosphate and cytochrome c. It is at this stage that cytochrome c oxidase is incorporated into the bilayer. In the final stage of dialysis the bilayer sheets fragment into small unilamellar vesicles. 3. The orientation of membrane protein in the final vesicles appears to be determined by the effect of protein conformation on the initial curvature of the bilayer sheets during the fragmentation process.  相似文献   

19.
The tumor promoter 20-3H-phorbol 12,13-dibutyrate bound in a specific manner to particulate preparations from both whole mouse skin and mouse epidermis. The binding, which was comparable in both whole skin and epidermal preparations, occurred rapidly, was reversible upon addition of non-radioactive ligand and showed high affinity (KD = 2.4 × 10?8 M). The potencies of phorbol esters for inhibiting binding of 3H-PDBu corresponded to their biological and tumor-promoting activities: phorbol 12-myristate 13-acetate, KI = 0.74 nM; phorbol 12,13-didecanoate, KI = 16 nM; phorbol 12,13-dibenzoate, KI = 82 nM; mezerein, KI = 98 nM; phorbol 12,13-diacetate, KI = 3 μM; phorbol 12,13,20-triacetate, KI = 5.6 μM; phorbol 13-acetate, KI = 64 μM. The biologically inactive derivatives phorbol (0.88 mM) and 4α-phorbol 12,13-didecanoate (15 μM) did not inhibit binding. Likewise, 3H-PDBu binding was only weakly inhibited by phorbol-related diterpenes which are highly inflammatory but nonpromoting. These structure-activity relationships suggest that the 3H-PDBu binding activity mediates phorbol ester tumor promotion. 3H-PDBu binding was not inhibited by the nonphorbol promoters examined. Similarly, it was not blocked by compounds which antagonize (dexamethasone acetate, 2 μM; retinoic acid, 10 μM) or mimic (epidermal growth factor, 100 ng/ml; melittin, 25 μg/ml; PGE2, 1 μM) some of the effects of the phorbol esters in vivo or in vitro.  相似文献   

20.
O2 uptake in spinach thylakoids was composed of ferredoxin-dependent and -independent components. The ferredoxin-independent component was largely 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU) insensitive (60%). Light-dependent O2 uptake was stimulated 7-fold by 70 μM ferredoxin and both uptake and evolution (with O2 as the only electron acceptor) responded almost linearly to ferredoxin up to 40 μM. NADP+ reduction, however, was saturated by less than 20 μM ferredoxin. The affinity of O2 uptake for for O2 was highly dependent on ferredoxin concentration, with K12(O2) of less than 20 μM at 2 μM ferredoxin but greater than 60 μM O2 with 25 μM ferredoxin. O2 uptake could be suppressed up to 80% with saturating NADP+ and it approximated a competitive inhibitor of O2 uptake with a Ki of 8–15 μM. Electron transport in these thylakoids supported high rates of photophosphorylation with NADP+ (600 μmol ATP/mg Chl per h) or O2 (280 μmol/mg Chl per h) as electron acceptors, with ATP2e ratios of 1.15–1.55. Variation in ATP2e ratios with ferredoxin concentration and effects of antimycin A indicate that cyclic electron flow may also be occurring in this thylakoid system. Results are discussed with regard to photoreduction of O2 as a potential source of ATP in vivo.  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