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.
1. A spectrophotometric assay is described that enables the hydroxylation of p-coumaric acid to caffeic acid, catalysed by spinach-beet phenolase, to be followed continuously. 2. Initial-velocity and inhibitor studies indicate that the order of substrate addition is oxygen, p-coumaric acid and electron donor, with an irreversible step separating the binding of each substrate. 3. Caffeic acid is most likely to act as electron donor at the active site; other electron donors, such as ascorbic acid, NADH and dimethyltetrahydropteridine, function mainly to recycle cofactor amounts of caffeic acid. 4. A reaction scheme, consistent with these data, is proposed.  相似文献   

2.
1. Chloroplasts isolated from leaves of spinach-beet (Beta vulgaris L. ssp. vulgaris) do not catalyse the hydroxylation of p-coumaric acid in the dark unless a reductant (such as ascorbate, NADH or NADPH) is added. Superoxide dismutase has no effect on this reaction. 2. Illuminated chloroplasts catalyse the hydroxylation in the absence of added reductant. This reaction is completely inhibited by superoxide dismutase, but catalase has little effect. 3. Both hydroxylation in the light and hydroxylation in the dark in the presence of reductants are inhibited by diethyldithiocarbamate, EDTA, cyanide and 2-mercaptoethanol. 4. It is proposed that O-2- generated by illuminated chloroplasts is involved in the provision of a reductant to the enzyme phenolase.  相似文献   

3.
1. Under defined conditions, the hydroxylation of p-coumaric acid catalysed by a phenolase from leaves of spinach beet (Beta vulgaris L.) was observed to develop its maximum rate only after a lag period. 2. By decreasing the reaction rate with lower enzyme concentrations or by increasing it with higher concentrations of reductants, the length of the lag period was inversely related to the maximum rate subsequently developed. 3. Low concentrations of caffeic acid or other o-dihydric phenols abolished this lag period. With caffeic acid, the rate of hydroxylation was independent of the reductant employed. 4. Hydroxylation was inhibited by diethyldithiocarbamate, but with low inhibitor concentrations hydroxylation recovered after a lag period. This lag could again be abolished by the addition of high concentrations of caffeic acid or other o-dihydric phenols. 5. Catechol oxidase activity showed no lag period, and did not recover from diethyldithiocarbamate inhibition. 6. The purified enzyme contained 0.17-0.33% copper; preparations with the highest specific activity were found to have the highest copper content. 7. The results are interpreted to suggest that the oxidation of o-dihydric phenols converts the enzymic copper into a species catalytically active in hydroxylation. This may represent the primary function for the catechol oxidase activity of the phenolase complex. The electron donors are concerned mainly, but not entirely, in the reduction of o-quinones produced in this reaction.  相似文献   

4.
1. In the presence of dihydroxyfumarate, horseradish peroxidase catalyses the conversion of p-coumaric acid into caffeic acid at pH 6. This hydroxylation is completely inhibited by superoxide dismutase. 2. Dihydroxyfumarate cannot be replaced by ascorbate H2O2, NADH, cysteine or sulphite. Peroxidase can be replaced by high (10 mM) concentrations of FeSO4, but this reaction is almost unaffected by superoxide dismutase. 3. Hydroxylation by the peroxidase/dihydroxyfumarate system is completely inhibited by low concentrations of Mn2+ or Cu2+. It is proposed that this is due to the ability of these metal ions to react with the superoxide radical O2--. 4. Hydroxylation is partially inhibited by mannitol, Tris or ethanol and completely inhibited by formate. This seems to be due to the ability of these reagents to react with the hydroxyl radical -OH. 5. It is concluded that O2-- is generated during the oxidation of dihydroxyfumarate by peroxidase and reacts with H2O2 to produce hydroxyl radicals, which then convert p-coumaric acid into caffeic acid.  相似文献   

5.
The effect of hydroxylation genes on the hydroxylation of intermediates of flavonoid biosynthesis in Petunia hybrida is reported. In mutants homozygous recessive, for the gene An9, dihydroflavonols accumulate. The number of hydroxyl groups in the B-ring is determined by the hydroxylation genes Htl and Hfl. A similar effect of Htl and (probably) Hfl occurs in flavanone-accumulating mutants, homozygous recessive for the gene An3. Mutants dominant for Hfl probably accumulate a 5,7,3,4,5-pentahydroxyflavanone. The mutant W43, homozygous recessive for the gene An5, is blocked in an early flavonoid biosynthesis step. It accumulates p-coumaric acid together with caffeic acid. The hydroxylation genes Htl and Hfl, however, are also homozygous recessive, which indicates that the hydroxylation of p-coumaric acid to caffeic acid or derivatives of these compounds is not controlled by Htl. The accumulation of caffeic acid was observed in all mutants investigated so far, regardless of which hydroxylation genes were dominant or recessive. We conclude that hydroxylations involved in anthocyanin biosynthesis occur at the C15 level.Deceased  相似文献   

6.
We previously have described a substance present in crude sonicates of L-929 cells which replaced ascorbate in vitro as a reductant for prolyl hydroxylase (B. Peterkofsky, D. Kalwinksy and R. Assad, 1980, Arch. Biochem. Biophys.199, 362–373). In the present study we found that almost 90% of the substance was particulate after differential centrifugation of stationary phase L-929 cell homogenates. The substance was not localized in nuclei or mitochondria and was found in the same fractions as microsomes, but these fractions also contained lysosomes and cell membranes. The reductant could not be solubilized from particles by Brij-35, indicating that it is an intrinsic component of a membrane rather than intracisternally located. The intramembranous cofactor, in the absence of ascorbate, participated in the in vitro hydroxylation of [4-3H]proline in radio-actively labeled, intracisternal unhydroxylated procollagen in isolated microsomes which also contained prolyl hydroxylase. Hydroxylation was determined by measuring tritiated water formed from release of the 4-trans tritium atom. Since it is unlikely that such participation could occur if the cofactor were located within the membrane of another subcellular organelle, we have concluded that it is in the same particle as prolyl hydroxylase and unhydroxylated procollagen, that is, the microsome. With the endogenous reductant the reaction was slower than with saturating ascorbate and was increased by NADH. Maximum hydroxylation with the endogenous reductant was close to that which could be achieved with ascorbate. These results provide strong evidence that the endogenous reductant alone can account for the phenomenon of ascorbate-independent proline hydroxylation in L-929 cells. As in the case of ascorbate, the microsomal reductant functioned only in the presence of α-ketoglutarate and Fe2+ and served as reductant for lysyl hydroxylase. It also was detected in the particulate fraction of virally transformed BALB 3T3 cells and in purified microsomes from bones of intact chick embryos. Since ascorbate could be taken up and concentrated in bone microsomes, it is unlikely that the endogenous reductant serves as an intermediary between ascorbate and intracisternal prolyl hydroxylase.  相似文献   

7.
Tyrosinase (EC 1.14.18.1) is a widely distributed type 3 copper enzyme participating in essential biological functions. Tyrosinases are potential biotools as biosensors or protein crosslinkers. Understanding the reaction mechanism of tyrosinases is fundamental for developing tyrosinase-based applications. The reaction mechanisms of tyrosinases from Trichoderma reesei (TrT) and Agaricus bisporus (AbT) were analyzed using three diphenolic substrates: caffeic acid, L-DOPA (3,4-dihydroxy-l-phenylalanine), and catechol. With caffeic acid the oxidation rates of TrT and AbT were comparable; whereas with L-DOPA or catechol a fast decrease in the oxidation rates was observed in the TrT-catalyzed reactions only, suggesting end product inhibition of TrT. Dopachrome was the only reaction end product formed by TrT- or AbT-catalyzed oxidation of L-DOPA. We produced dopachrome by AbT-catalyzed oxidation of L-DOPA and analyzed the TrT end product (i.e. dopachrome) inhibition by oxygen consumption measurement. In the presence of 1.5mM dopachrome the oxygen consumption rate of TrT on 8mM L-DOPA was halved. The type of inhibition of potential inhibitors for TrT was studied using p-coumaric acid (monophenol) and caffeic acid (diphenol) as substrates. The strongest inhibitors were potassium cyanide for the TrT-monophenolase activity, and kojic acid for the TrT-diphenolase activity. The lag period related to the TrT-catalyzed oxidation of monophenol was prolonged by kojic acid, sodium azide and arbutin; contrary it was reduced by potassium cyanide. Furthermore, sodium azide slowed down the initial oxidation rate of TrT- and AbT-catalyzed oxidation of L-DOPA or catechol, but it also formed adducts with the reaction end products, i.e., dopachrome and o-benzoquinone.  相似文献   

8.
Microsomes from chronic ethanol-fed rats were previously shown to catalyze the NADPH-dependent production of reactive oxygen intermediates at elevated rates compared to controls. Recent studies have shown that NADH can also serve as a reductant and promote the production of oxygen radicals by microsomes. The current study evaluated the influence of chronic ethanol consumption on NADH-dependent microsomal production of reactive oxygen intermediates, and compared the results with NADH to those of NADPH. Microsomal oxidation of chemical scavengers, taken as a reflection of the production of hydroxyl radical (.OH)-like species was increased about 50% with NADH as cofactor and about 100% with NADPH after chronic ethanol consumption. The potent inhibition of the production of .OH-like species by catalase suggests a precursor role for H2O2 in .OH production. Rates of NADH- and NADPH-dependent H2O2 production were increased by about 50 and 70%, respectively, after chronic ethanol consumption. A close correlation between rates of H2O2 production and generation of .OH-like species was observed for both NADH and NADPH, and increased rates of H2O2 production appear to play an important role in the elevated generation of .OH-like species after chronic ethanol treatment. Microsomal lipid peroxidation was elevated about 60% with NADH, and 120% with NADPH, after ethanol feeding. With both types of microsomal preparations, the characteristics of the NADH-dependent reactions were similar to the NADPH-dependent reactions, e.g., sensitivity to antioxidants and free radical scavengers and catalytic effectiveness of ferric complexes. However, rates with NADPH exceeded the NADH-dependent rates by 50 to 100%, and the increased production of reactive oxygen intermediates by microsomes after ethanol treatment was greater with NADPH (about twofold) than with NADH (about 50%). Oxidation of ethanol results in an increase in hepatic NADH levels and interaction of NADH, iron, and microsomes can produce potent oxidants capable of initiating lipid peroxidation and oxidizing .OH scavengers. These acute metabolic interactions produced by ethanol-derived NADH are increased, not attenuated, in microsomes from chronic ethanol-fed rats, and it is possible that such increases in NADH (and NADPH)-dependent production of reactive oxygen species play a role in the development of oxidative stress in the liver as a consequence of ethanol treatment.  相似文献   

9.
The uncoupling of 2-oxoglutarate decarboxylation from hydroxylation in the reaction catalyzed by thymine 7-hydroxylase (thymine, 2-oxoglutarate:oxygen oxidoreductase (7-hydroxylating), EC 1.14.11.6) in the presence of 5-fluorouracil has been studied. In the complete reaction no external reductant is formally needed. The uncoupled reaction is almost negligible in the absence of ascorbate and the optimal ascorbate concentration is 5-times higher than in the presence of a hydroxylatable substrate. This indicates that ascorbate acts as the external reductant that is formally needed in the catalytic cycle. The complete reaction follows the steady-state kinetics of an ordered ter reactant mechanism where 2-oxoglutarate and thymine have to be bound to the enzyme before oxygen (E. Holme (1975) Biochemistry 14, 4999-5003). The uncoupled reaction follows the same kinetic pattern as the complete reaction, and in accordance with this no decarboxylation of 2-oxoglutarate occurs in the absence of a substrate analogue even at elevated oxygen tension. There is a good agreement between Kia values for 2-oxoglutarate of the two reactions, but there is at least a 6-fold increase in KO2 where a minimum value of 25% O2 in the gas phase was found for the partial reaction. The high KO2 found means that the reaction rate could increase considerably at elevated oxygen tension.  相似文献   

10.
Treatment of spinach-beet phenolase with H2O2 under aerobic conditions results in a stimulation of the p-coumaric acid hydroxylation it catalyses, but not the caffeic acid oxidation. Spectroscopic evidence suggests that an oxygenated enzyme species is formed under these conditions.  相似文献   

11.
Microsomal preparations from parsley cell suspension cultures challenged with an elicitor from Phytophthora megasperma f.sp. glycinea (Pmg) catalyze the formation of trans-5-O-caffeoylshikimate from trans-5-O-(4-coumaroyl)shikimate. Neither the cis isomer nor free 4-coumarate, 4-coumaroyl-CoA, or 5-O-(4-coumaroyl)quinate are substrates for this enzyme. The reaction is strictly dependent on NADPH as a reducing cofactor and on molecular oxygen. NADH, ascorbic acid, and 6,7-dimethyl-5,6,7,8-tetrahydropterine cannot substitute for NADPH. However, NADH enhances enzyme activity observed in the presence of NADPH. Cytochrome c and carbon monoxide inhibit the hydroxylation reaction, suggesting a cytochrome P-450-dependent mixed-function monooxygenase.  相似文献   

12.
p-Coumaroyl-D-glucose hydroxylase in sweet potato (Ipomoea batatas Lam.) has been purified to apparent electrophoretic homogeneity using a combination of anion-and cation-exchange, hydrophobic and gel filtration chromatography. The purified enzyme was a monomer with a molecular weight of 33,000 and pI of 8.3. The purified enzyme showed not only hydroxylase activity but also polyphenol oxidase activity. L-Ascorbic acid was the best electron donor for the hydroxylation reaction, which had an optimum pH of 7.0. The enzyme hydroxylated p-coumaroyl-D-glucose, p-coumaric acid, and p-cresol but did not act on o-coumaric acid, m-coumaric acid, 4-hydroxy-3-methoxycinnamic acid, p-hydroxybenzoic acid or L-tyrosine. While the enzyme utilized p-coumaroyl-D-glucose and p-coumaric acid equally at pH 7.0, it hydroxylated only p-coumaroyl-D-glucose at pH 5.5. The enzyme oxidized diphenols such as D,L-(3,4-dihydroxyphenyl) alanine and caffeic acid, but exhibited no clear pH optimum in this reaction characteristic of polyphenol oxidase. Both the hydroxylase and the polyphenol oxidase activities were strongly inhibited by beta-mercaptoethanol, diethyldithiocarbamate, KCN, and p-coumaric acid (in concentrations higher than 5 mM). Ammonium sulfate and sodium chloride activated the hydroxylase activity but not the polyphenol oxidase activity of the enzyme. The enzyme activity and L-ascorbic acid contents changed in a manner suggesting their involvements in chlorogenic acid biosynthesis during incubation of sliced sweet potato root tissues.  相似文献   

13.
Reductant used as cofactor for the prolyl hydroxylase reaction, was measured by a tritium release assay modified from an enzyme assay by making all components of the assay system saturating except for the reductant, but including prolyl hydroxylase. Reduced glutathione (6 mm), which had little activity as a cofactor, and thymol (0.1 mm), an antioxidant which exhibited no cofactor activity at all, were required for optimal proline hydroxylation dependent on reducing cofactor, with thymol fulfilling the previously described requirement for catalase. Ascorbate, cysteine and 6,7-dimethyltetrahydropterin were active as cofactors, in descending order of activity at equimolar concentrations, and activity was concentration dependent for all of these compounds. Sonicates of stationary phase L-929 cells which exhibit ascorbate-independent proline hydroxylation in culture contained reducing cofactor which could replace ascorbate in the cofactor assay, while sonicates of log phase cells which exhibit an ascorbate requirement in culture contained about one-third or less of that amount. NADH and NADPH, which themselves have little or no activity as cofactor, increased the cofactor activity of log phase cell sonicates but had relatively little effect on the activity of stationary cell sonicates suggesting that the cofactor is in a more reduced state in stationary phase. Within 24 h after replating dense, stationary phase cell cultures at low density, conditions where cells return to ascorbate dependence, prolyl hydroxylase activity had decreased to one-fifth the original activity while the concentration of functional reducing cofactor had decreased to less than 1% of its original concentration, largely as a result of oxidation. Ascorbate was not present in L-929 cells sonicates and the levels of tetrahydropterin and cysteine in sonicates could not account for the amount of cofactor activity exhibited by the sonicates in the assay system. Treatment of L-929 cultures with aminopterin did not decrease ascorbate independence, suggesting that tetrahydrofolate did not contribute significantly to cellular proline hydroxylation. These results suggest that an unidentified reductant present in L-929 cells can account for ascorbate-independent proline hydroxylation and also regulate prolyl hydroxylase activity in these cells and that cellular levels of reduced pyridine nucleotides may regulate the reduction state of this substance.  相似文献   

14.
Lineweaver-Burk plots of reduced nicotinamide adenine dinucleotide (NADH) oxidation by membrane preparations from Bacillus subtilis are biphasic, with two K(m) values for NADH. The higher K(m) corresponds to the only K(m) observed for NADH oxidation by whole cells, whereas the lower K(m) corresponds to that observed with open cell envelopes. Membrane preparations apparently contain a small fraction of open or inverted vesicles which is responsible for the low K(m) reaction, whereas entry of NADH into the larger portion of closed, normally oriented vesicles is rate limiting and responsible for the high K(m) reaction. In contrast, the oxidation of l-alpha-glycerol-phosphate (glycerol-P) by membrane preparations shows only one K(m) that corresponds to that of glycerol-P oxidation by whole cells or lysates. Since glycerol-P dehydrogenase (NAD independent) has the same K(m), this enzyme reaction rather than entry of glycerol-P into vesicles represents the rate-limiting step for glycerol-phosphate oxidation. The K(m) for amino acid uptake by vesicles in the presence of NADH corresponds to the high K(m) for NADH oxidation, indicating that NADH energizes transport only if it enters closed, normally oriented vesicles. Studies with rotenone and proteolytic enzymes support this interpretation. The apparent efficiency of NADH in energizing uptake seems to be lower than that of glycerol-P because, under the experimental conditions usually employed, open or inverted vesicles that do not participate in amino acid uptake are responsible for the major portion of NADH oxidation. When the results are corrected for this effect, the efficiency of NADH is essentially the same as that of l-alpha-glycerol-P.  相似文献   

15.
Acetate and other short chain n-fatty acids (C(1)-C(6)) inhibit strongly the uptake of l-serine or other l-amino acids but inhibit only weakly that of alpha-methylglucoside or fructose, whether measured in whole cells of Bacillus subtilis or in membrane vesicles that have been energized with reduced nicotinamide adenine dinucleotide (NADH), l-alpha-glycerol phosphate, or ascorbate plus phenazine methosulfate. The acetate inhibition is noncompetitive, as was shown for l-alpha-aminoisobutyric acid uptake by whole cells and for l-serine uptake by membrane vesicles. In membrane preparations, neither NADH oxidation nor the reduction of cytochromes by NADH are affected by fatty acids. All of these effects are similar to those of 2, 4-dinitrophenol. It is concluded that the fatty acids "uncouple" the amino acid carrier proteins from the cytochrome-linked electron transport system (to which they may be coupled via protein interaction or via a cation gradient).  相似文献   

16.
This study addresses the dynamic interactions among alpha-tocopherol, caffeic acid, and ascorbate in terms of a sequence of redox cycles aimed at accomplishing optimal synergistic antioxidant protection. Several experimental models were designed to examine these interactions: UV irradiation of alpha-tocopherol-containing sodium dodecyl sulfate micelles, one-electron oxidations catalyzed by the hypervalent state of myoglobin, ferrylmyoglobin, and autoxidation at appropriate pHs. These models were assessed by ultraviolet (UV) and electron paramagnetic resonance (EPR), entailing direct- and continuous-flow experiments, spectroscopy and by separation and identification of products by HPLC. The alpha-tocopheroxyl radical EPR signal generated by UV irradiation of alpha-tocopherol-containing micelles was suppressed by caffeic acid and ascorbate; in the former case, no other EPR signal was observed at pH 7.4, whereas in the latter case, the alpha-tocopheroxyl radical EPR signal was replaced by a doublet EPR spectrum corresponding to the ascorbyl radical (A*-). The potential interactions between caffeic acid and ascorbate were further analyzed by assessing, on the one hand, the ability of ascorbate to reduce the caffeic acid o-semiquinone (generated by oxidation of caffeic acid by ferrylmyoglobin) and, on the other hand, the ability of caffeic acid to reduce ascorbyl radical (generated by autoxidation or oxidation of ascorbate by ferrylmyoglobin). The data presented indicate that the reductive decay of ascorbyl radical (A*-) and caffeic acid o-semiquinone (Caf-O*) can be accomplished by caffeic acid (Caf-OH) and ascorbate (AH-), respectively, thus pointing to the reversibility of the reaction Caf-O* + AH- <--> Caf-OH + A*-. Continuous-flow EPR measurements of mixtures containing ferrylmyoglobin, alpha-tocopherol-containing micelles, caffeic acid, and ascorbate revealed that ascorbate is the ultimate electron donor in the sequence encompassing transfer of the radical character from the micellar phase to the phase. In independent experiments, the effects of caffeic acid and ascorbate on the oxidation of two low-density lipoprotein (LDL) populations, control and alpha-tocopherol-enriched, were studied and results indicated that alpha-tocopherol, caffeic acid, and ascorbate acted synergistically to afford optimal protection of LDL against oxidation. These results are analyzed for each individual antioxidant in terms of three domains: its localization and that of the antioxidant-derived radical, its reduction potential, and the predominant decay pathways for the antioxidant-derived radical, that exert kinetic control on the process.  相似文献   

17.
Microsomal preparations from carrot (Daucus carota L.) cell suspension cultures catalyze the formation of trans-5-O-caffeoyl-D-quinate (chlorogenate) from trans-5-O-(4-coumaroyl)-D-quinate. trans-5-O-(4-Coumaroyl)shikimate is converted to about the same extent to trans-5-O-caffeoylshikimate. trans-4-O-(4-Coumaroyl)-D-quinate, trans-3-O-(4-coumaroyl)-D-quinate, trans-4-coumarate, and cis-5-O-(4-coumaroyl)-D-quinate do not act as substrates. The reaction is strictly dependent on molecular oxygen and on NADPH as reducing cofactor. NADH and ascorbic acid cannot substitute for NADPH. Cytochrome c, Tetcyclacis, and carbon monoxide inhibit the reaction suggesting a cytochrome P-450-dependent mixed-function monooxygenase. Competition experiments as well as induction and inhibition phenomena indicate that there is only one enzyme species which is responsibl for the hydroxylation of the 5-O-(4-coumaric) esters of both D-quinate and shikimate. The activity of this enzyme is greatly increased by in vivo irradiation of the cells with blue/uv light. We conclude that the biosynthesis of the predominant caffeic acid conjugates in carrot cells occurs via the corresponding 4-coumaric acid esters. Thus, in this system, 5-O-(4-coumaroyl)-D-quinate can be seen as the final intermediate in the chlorogenic acid pathway.  相似文献   

18.
The 4- and 5-hydroxylations of phenolic compounds in plants are catalyzed by cytochrome P450 enzymes. The 3-hydroxylation step leading to the formation of caffeic acid from p-coumaric acid remained elusive, however, alternatively described as a phenol oxidase, a dioxygenase, or a P450 enzyme, with no decisive evidence for the involvement of any in the reaction in planta. In this study, we show that the gene encoding CYP98A3, which was the best possible P450 candidate for a 3-hydroxylase in the Arabidopsis genome, is highly expressed in inflorescence stems and wounded tissues. Recombinant CYP98A3 expressed in yeast did not metabolize free p-coumaric acid or its glucose or CoA esters, p-coumaraldehyde, or p-coumaryl alcohol, but very actively converted the 5-O-shikimate and 5-O-d-quinate esters of trans-p-coumaric acid into the corresponding caffeic acid conjugates. The shikimate ester was converted four times faster than the quinate derivative. Antibodies directed against recombinant CYP98A3 specifically revealed differentiating vascular tissues in stem and root. Taken together, these data show that CYP98A3 catalyzes the synthesis of chlorogenic acid and very likely also the 3-hydroxylation of lignin monomers. This hydroxylation occurs on depsides, the function of which was so far not understood, revealing an additional and unexpected level of networking in lignin biosynthesis.  相似文献   

19.
In spinach thylakoids prepared from intact chloroplasts by shocking in the presence of ascorbate to preserve the operation of ascorbate peroxidase, the rate of oxygen uptake with methyl viologen as acceptor decreased in response to the addition of H2O2. Such a decrease was not observed in the presence of KCN or when the thylakoids lost ascorbate peroxidase activity. Illumination of intact chloroplasts in the presence of H2O2 and methyl viologen showed an initial rate of oxygen exchange, which is intermediate between the initial rate of oxygen evolution in the presence of H2O2 alone and steady-state oxygen uptake in the presence of methyl viologen. The data showed that monodehydroascorbate radical generated in ascorbate peroxidase reaction could compete with methyl viologen for electrons supplied by the electron transport chain in both thylakoids and intact chloroplasts. During the illumination of intact chloroplasts the rate of oxygen uptake increased. The presence of nigericin swiftly led to steady-state oxygen uptake, and to a clear-cut 1:1 relationship between the electron transport rate estimated from fluorescence assay and the electron transport rate determined from oxygen uptake, taking the stoichiometry 1O2:4e. The increase in oxygen uptake was attributed to the cessation of monodehydroascorbate radical generation brought about by consumption of intrachloroplast ascorbate in the peroxidase reactions, and the effects of nigericin were explained by acceleration of such consumption. The competition between methyl viologen and monodehydroascorbate radical in the intact chloroplasts was estimated under various conditions.  相似文献   

20.
Externally added ascorbate or NADH effectively reduced ferricyanide and promoted the exit of Fe3+ originated from acid-destabilized transferrin contained inside endocytic vesicles. The effect of ascorbate was mediated by an ascorbate uptake system, and the effect of NADH was mediated by the membrane-associated oxidoreductase. At physiological concentrations of both ascorbate and NADH, the ascorbate transport and the NADH-oxidoreductase system were additive as measured by the rate of reduction of ferricyanide and by the mobilization of transferrin-associated iron. The results indicate that Fe3+ reduction may occur by a nonenzymatic reaction with ascorbate transported into the vesicle lumen. The ascorbate-mediated reduction of iron derived from transferrin occurring in the endosome could play a major role in cellular iron uptake.  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