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.
I. L. Sun  E. E. Sun  F. L. Crane 《Protoplasma》1995,184(1-4):214-219
Summary The addition of coenzyme Q10 to culture media stimulates the serum-free growth of HeLa, HL-60 cells, and mouse fibroblasts (Balb/3T3). With HeLa cells, the stimulation by coenzyme Q10 is additive to the stimulation by ferricyanide, an impermeable electron acceptor for the transplasma membrane electron transport. This combined response to coenzyme Q10 and ferricyanide is enhanced with insulin. -Tocopherylquinone can also stimulate the growth of HeLa cells, but vitamin K1 is inactive. Specificity of quinone effects is indicated. Serum-free growth of Balb/3T3 and SV 40 transformed BaIb/3T3 (SV/T2) cells is also stimulated by coenzyme Qio with stimulation similar to HeLa cells. However, Balb/3T3 cells are not stimulated by ferricyanide, which does not increase the response to coenzyme Q10. The transformed cells (SV/T2) respond better to ferricyanide alone, but the effects of coenzyme Qio and ferricyanide are not additive. Serum-free growth of HL-60 cells is stimulated dramatically by coenzyme Q10. The extent of growth stimulation on HL-60 cells is almost six-fold that of HeLa or Balb/3T3 cells. The stimulation of NADH-ferricyanide reductase (a transmembrane redox enzyme) by coenzyme Q10 with HL-60 cells is similar to their growth pattern in response to coenzyme Q10. Unlike HL-60, HeLa and Balb/3T3 cells show little stimulation of ferricyanide reduction by coenzyme Q10. The stimulatory effect on both ferricyanide reduction and cell growth by the short side-chain coenzyme Q2 is much less than that of the long side-chain coenzyme Q10. Ferricyanide reduction by HeLa cells is inhibited by coenzyme Q analogs such as 2,3-dimethoxy-5-chloro-6-naphthyl-mercapto-coenzyme Q and 2-methoxy-3-ethoxyl-5-methyl-6-hexadecyl-mercapto-coenzyme Q. However, these inhibitions are reversed by coenzyme Q10. The growth inhibition of HL-60 cells by other coenzyme Q analogs, such as capsiacin can also be reversed by coenzyme Q10. These data indicate that plasma membrane-based NADH oxidation or modification of the membrane quinone redox balance may be a basis for the growth stimulation.  相似文献   

2.
A role for coenzyme Q in the stabilization of extracellular ascorbate by intact cells has beenrecently recognized. The aim of this work was to study the interactions between reducedubiquinone in the plasma membrane and the ascorbyl free radical, as an approach to understandubiquinone-mediated ascorbate stabilization at the cell surface. K-562 cells stabilized ascorbateand decreased the steady-state levels of the semiascorbyl radical. The ability of cells to reduceascorbyl free radical was inhibited by the quinone analogs capsaicin and chloroquine andstimulated by supplementing cells with coenzyme Q10. Purified plasma membranes also reducedascorbyl free radical in the presence of NADH. Free-radical reduction was notobserved inquinone-depleted plasma membranes, but restored after its reconstitution with coenzyme Q10.Addition of reduced coenzyme Q10 to depleted membranes allowed them toreduce the signalof the ascorbyl free radical without NADH incubation and the addition of an extra amount ofpurified plasma membrane quinone reductase further stimulated this activity. Reduction wasabolished by treatment with the reductase inhibitor p-hydroximercuribenzoate and by blockingsurface glycoconjugates with the lectin wheat germ agglutinin, which supports the participationof transmembrane electron flow. The activity showed saturation kinetics by NADH andcoenzyme Q, but not by the ascorbyl free radical in the range of concentrations used. Our resultssupport that reduction of ascorbyl free radicals at the cell surface involves coenzyme Qreduction by NADH and the membrane-mediated reduction of ascorbyl free radical.  相似文献   

3.
Rate constants for reduction of cytochrome b561 by internal ascorbate (k0A) and oxidation by external ferricyanide (k1F) were determined as a function of pH from rates of steady-state electron transfer across chromaffin-vesicle membranes. The pH dependence of electron transfer from cytochrome b561 to ferricyanide (k1F) may be attributed to the pH dependence of the membrane surface potential. The rate constant for reduction by internal ascorbate (k0A), like the previously measured rate constant for reduction by external ascorbate (k-1A), is not very pH-dependent and is not consistent with reduction of cytochrome b561 by the ascorbate dianion. The rate at which ascorbate reduces cytochrome b561 is orders of magnitude faster than the rate at which it reduces cytochrome c, despite the fact that midpoint reduction potentials favor reduction of cytochrome c. Moreover, the rate constant for oxidation of cytochrome b561 by ferricyanide (k1F) is smaller than the previously measured rate constant for oxidation by semidehydroascorbate, despite the fact that ferricyanide has a higher midpoint reduction potential. These results may be reconciled by a mechanism in which electron transfer between cytochrome b561 and ascorbate/semidehydroascorbate is accelerated by concerted transfer of a proton. This may be a general property of biologically significant electron transfer reactions of ascorbic acid.  相似文献   

4.
Ascorbate was maintained in the media during a long-term culture by HL-60 cells. The chemical oxidation of ascorbate was reversed in vitro by living HL-60 cells and was related to the amount of cells added. The increase of NADH concentration by lactate addition to cells was accompanied by an increase of both ascorbate regeneration and ferricyanide reduction. Further, plasma membrane enriched fractions from HL-60 cells revealed enhancement of both ascorbate regeneration and ferricyanide reduction in the presence of NADH when previously treated with detergent. The blockage of cell surface carbohydrates by wheat germ agglutinin (WGA) and Concanavalina ensiformis (Con A) lectins significantly inhibited the regeneration of ascorbate caused by the cells. These results support the idea that ascorbate is externally regenerated by the NADH-ascorbate free radical reductase as a part of the transplasma membrane redox system.  相似文献   

5.
Ascorbate free radical and its role in growth control   总被引:4,自引:0,他引:4  
Summary Ascorbate and its free radical potentiates proliferation of HL-60 cells in serum-limiting media. Dehydroascorbate does not affect growth. This stimulation of growth is due to a general shortening of the cell cycle. The incubation of HL-60 cells with ascorbate free radical produces a significant change of the redox potential of cells. The presence of cells in culture media avoids the total oxidation of ascorbate, and also HL-60 cells induce the short-term stabilization of ascorbate. Ascorbate free radical potentiates also the onset of DNA synthesis in CCL39 cells induced by fetal calf serum, although itself does not affect quiescense to proliferation transition. This transition induced by fetal calf serum also potentiates the capacity of CCL39 cells to stabilize ascorbate. We discuss here the role of ascorbate free radical on growth control by its reduction by the plasma membrane redox system and its meaning for cell physioslogy.  相似文献   

6.
The glyceollin inhibition of electron transport by isolated soybean and corn mitochondria was similar to that of rotenone, acting at site I between the internal NADH dehydrogenase and coenzyme Q. Coupled state 3 malate oxidation was inhibited by glyceollin and rotenone with apparent Ki values of about 15 and 5 micromolar, respectively. Carbonylcyanide m-chlorophenyl hydrazone uncoupled state 4 malate oxidation was also inhibited by glyceollin and rotenone, but uncoupled succinate and exogenous NADH state 4 oxidation was only slightly inhibited by both compounds. Glyceollin also inhibited ferricyanide reduction with malate as the electron donor, with an apparent Ki of 5.4 micromolar, but failed to inhibit such reduction with succinate or externally added NADH as electron donors. Glyceollin did not inhibit state 4 oxidation of malate, succinate, or exogenous NADH. Glyceollin did not act as a classical uncoupler or as an inhibitor of oxidative phosphorylation.  相似文献   

7.
Stable nitroxide radicals have been considered as therapeutic antioxidants because they can scavenge more toxic radicals in biologic systems. However, as radicals they also have the potential to increase oxidant stress in cells and tissues. We studied the extent to which this occurs in cultured EA.hy926 endothelial cells exposed to the nitroxide Tempol (4-hydroxy-2,2,6,6-tetramethylpiperidine-N-oxyl). Tempol was rapidly reduced by the cells, as manifest by an increase in the ability of the cells to reduce extracellular ferricyanide and by disappearance of the Tempol EPR signal. Cells loaded with ascorbic acid, which directly reacts with Tempol, showed increased rates of Tempol-dependent ferricyanide reduction, and a more rapid loss of the Tempol EPR signal than cells not containing ascorbate. In this process, intracellular ascorbate was oxidized, and was depleted at lower Tempol concentrations than was GSH, another important intracellular low molecular weight antioxidant. Further evidence that Tempol concentrations of 100-1000 μM induced an oxidant stress was that it caused an increase in the oxidation of dihydrofluorescein in cells and inhibited ascorbate transport at concentrations as low as 50-100 μM. The presence of intracellular ascorbate both prevented dihydrofluorescein oxidation and spared GSH from oxidation by Tempol. Such sparing was not observed when GSH was depleted by other mechanisms, indicating that it was likely due to protection against oxidant stress. These results show that whereas Tempol may scavenge other more toxic radicals, care must be taken to ensure that it does not itself induce an oxidant stress, especially with regard to depletion of ascorbic acid.  相似文献   

8.
Plasma membranes isolated from wild-type Saccharomyces cerevisiae crude membrane fractions catalyzed NADH oxidation using a variety of electron acceptors, such as ferricyanide, cytochrome c, and ascorbate free radical. Plasma membranes from the deletion mutant strain coq3, defective in coenzyme Q (ubiquinone) biosynthesis, were completely devoid of coenzyme Q6 and contained greatly diminished levels of NADH–ascorbate free radical reductase activity (about 10% of wild-type yeasts). In contrast, the lack of coenzyme Q6 in these membranes resulted in only a partial inhibition of either the ferricyanide or cytochrome-c reductase. Coenzyme Q dependence of ferricyanide and cytochrome-c reductases was based mainly on superoxide generation by one-electron reduction of quinones to semiquinones. Ascorbate free radical reductase was unique because it was highly dependent on coenzyme Q and did not involve superoxide since it was not affected by superoxide dismutase (SOD). Both coenzyme Q6 and NADH–ascorbate free radical reductase were rescued in plasma membranes derived from a strain obtained by transformation of the coq3 strain with a single-copy plasmid bearing the wild type COQ3 gene and in plasma membranes isolated form the coq3 strain grown in the presence of coenzyme Q6. The enzyme activity was inhibited by the quinone antagonists chloroquine and dicumarol, and after membrane solubilization with the nondenaturing detergent Zwittergent 3–14. The various inhibitors used did not affect residual ascorbate free radical reductase of the coq3 strain. Ascorbate free radical reductase was not altered significantly in mutants atp2 and cor1 which are also respiration-deficient but not defective in ubiquinone biosynthesis, demonstrating that the lack of ascorbate free radical reductase in coq3 mutants is related solely to the inability to synthesize ubiquinone and not to the respiratory-defective phenotype. For the first time, our results provide genetic evidence for the participation of ubiquinone in NADH–ascorbate free radical reductase, as a source of electrons for transmembrane ascorbate stabilization.  相似文献   

9.
Summary Ascorbate free radical is considered to be a substrate for a plasma membrane redox system in eukaryotic cells. Moreover, it might be involved in stimulation of cell proliferation. Ascorbate free radical can be generated by autoxidation of the ascorbate dianion, by transition metal-dependent oxidation of ascorbate, or by an equilibrium reaction of ascorbate with dehydroascorbic acid. In this study, we investigated the formation of ascorbate free radical, at physiological pH, in mixtures of ascorbate and dehydroascorbic acid by electron spin resonance spectroscopy. It was found that at ascorbate concentrations lower than 2.5 mM, ascorbate-free radical formation was not dependent on the presence of dehydroascorbic acid. Removal of metal ions by treatment with Chelex 100 showed that autoxidation under these conditions was less than 20%. Therefore, it is concluded that at low ascorbate concentrations generation of ascorbate free radical mainly proceeds through metal-ion-dependent reactions. When ascorbate was present at concentrations higher than 2.5 mM, the presence of dehydroascorbic acid increased the ascorbate free-radical signal intensity. This indicates that under these conditions ascorbate free radical is formed by a disproportionation reaction between ascorbate and dehydroascorbic acid, having aK equil of 6 × 10–17 M. Finally, it was found that the presence of excess ferricyanide completely abolished ascorbate free-radical signals, and that the reaction between ascorbate and ferricyanide yields dehydroascorbic acid. We conclude that, for studies under physiological conditions, ascorbate free-radical concentrations cannot be calculated from the disproportionation reaction, but should be determined experimentally.Abbreviations AFR ascorbate free radical - DHA dehydroascorbic acid - EDTA ethylenediaminetetraacetic acid - DTPA diethylenetri-aminepentaacetic acid - TEMPO 2,2,6,6-tetramethylpiperidinoxy  相似文献   

10.
P. Horton  W. A. Cramer 《BBA》1974,368(3):348-360
(1) (a) A concentration range of ferricyanide ( 0.125–0.5 mM) can be found which in the dark causes oxidation of cytochrome ƒ with two distinct kinetic components of comparable amplitude. The slow oxidation has a half time of 1–2 min. (b) The oxidation of cytochrome ƒ by ferricyanide is rapid and monophasic after the chloroplasts are frozen and thawed. (c) The oxidation of cytochrome b-559 by ferricyanide in the dark is mostly monophasic with a time course similar to that of the fast component in the cytochrome ƒ oxidation. (d) Ascorbate reduction of cytochromes ƒ and b-559 appears monophasic. Reduction of cytochrome b-559 by ascorbate is somewhat faster, and that by hydroquinone somewhat slower, than the corresponding reduction of cytochrome ƒ.

(2) (a) The kinetics of dark ferricyanide oxidation of cytochrome ƒ after actinic preillumination in the presence of an electron acceptor are approximately monophasic with a half time of about 30 s and do not show the presence of the slowly oxidized component observed after prolonged dark incubation. (b) The effect of actinic preillumination in altering the time course of ferricyanide oxidation appears to persist for several minutes in the dark. (c) Preillumination causes an increase in the extent of cytochrome b-559 oxidation by low concentrations of ferricyanide. The increase is inhibited if 3-(3′,4′-dichlorophenyl)-1,1-dimethylurea is present during the preillumination. (d) The presence of 3-(3′,4′-dichlorophenyl)-1,1-dimethylurea during preillumination does not inhibit the amplitude or rate of ferricyanide oxidation of cytochrome ƒ, although the presence of the inhibitor KCN does cause such inhibition.

(3) It is proposed that a significant fraction of the cytochrome ƒ population resides at a position in the membrane relatively inaccessible to the aqueous interface compared to high potential cytochrome b-559. Actinic illumination would cause a structural or conformational change in the cytochrome ƒ and/or the membrane resulting in an increase in accessibility to this fraction of the cytochrome ƒ population.  相似文献   


11.
Plasma membranes isolated from K562 cells contain an NADH-ascorbate free radical reductase activity and intact cells show the capacity to reduce the rate of chemical oxidation of ascorbate leading to its stabilization at the extracellular space. Both activities are stimulated by CoQ10 and inhibited by capsaicin and dicumarol. A 34-kDa protein (p34) isolated from pig liver plasma membrane, displaying NADH-CoQ10 reductase activity and its internal sequence being identical to cytochrome b 5 reductase, increases the NADH-ascorbate free radical reductase activity of K562 cells plasma membranes. Also, the incorporation of this protein into K562 cells by p34-reconstituted liposomes also increased the stabilization of ascorbate by these cells. TPA-induced differentiation of K562 cells increases ascorbate stabilization by whole cells and both NADH-ascorbate free radical reductase and CoQ10 content in isolated plasma membranes. We show here the role of CoQ10 and its NADH-dependent reductase in both plasma membrane NADH-ascorbate free radical reductase and ascorbate stabilization by K562 cells. These data support the idea that besides intracellular cytochrome b 5-dependent ascorbate regeneration, the extracellular stabilization of ascorbate is mediated by CoQ10 and its NADH-dependent reductase.  相似文献   

12.
Evidence for coenzyme Q function in transplasma membrane electron transport   总被引:2,自引:0,他引:2  
Transplasma membrane electron transport activity has been associated with stimulation of cell growth. Coenzyme Q is present in plasma membranes and because of its lipid solubility would be a logical carrier to transport electrons across the plasma membrane. Extraction of coenzyme Q from isolated rat liver plasma membranes decreases the NADH ferricyanide reductase and added coenzyme Q10 restores the activity. Piericidin and other analogs of coenzyme Q inhibit transplasma membrane electron transport as measured by ferricyanide reduction by intact cells and NADH ferricyanide reduction by isolated plasma membranes. The inhibition by the analogs is reversed by added coenzyme Q10. Thus, coenzyme Q in plasma membrane may act as a transmembrane electron carrier for the redox system which has been shown to control cell growth.  相似文献   

13.
West J  Hill R 《Plant physiology》1967,42(6):819-826
Pea chloroplasts isolated in salt media show decreased rates of 2:6 dichlorophenolindophenol (DCPIP) and ferricyanide reduction when depleted of CO2 at pH values below 7.5. The greatest effect of CO2 was on uncoupled systems. The incorporation of 10−2, 2 × 10−2 and 4 × 10−2 m sodium acetate into the reaction mixtures progressively increased the bicarbonate concentration required for half maximal rates of reduction of DCPIP. The reaction was saturated by bicarbonate concentrations of 1 to 4 × 10−2 m. With both DCPIP and ferricyanide, the addition of bicarbonate to illuminated chloroplast systems depleted of CO2 gave very rapid increases in the rates of reduction. Bicarbonate also stimulated oxygen uptake by the illuminated chloroplasts when added hydrogen acceptors had been reduced. There was no effect of bicarbonate on ferricyanide reduction at low light intensities, but with DCPIP reduction, the apparent magnitude of the effect was independent of light intensity. This suggests that DCPIP reacts with the chloroplast electron transport chain at a site nearer to a photochemical stage than does ferricyanide. It also suggests that CO2 has at least 2 sites of action.  相似文献   

14.
The absorbance and EPR spectra of type 1 and 2 copper-binding centres which are present in ceruloplasmin (Cp) molecule were shown to disappear upon the reduction of the enzyme by ascorbate under anaerobic conditions. The fluorescence band attributed to type 3 Cu was altered concomitantly. The electron-accepting nitroxyl radical added to reduced Cp restored the absorbance, EPR and fluorescence spectra of the oxidase. Only type 1 and 3 copper ions, as judged by spectral changes, can be reduced by ascorbate and then reoxidized by the nitroxyl radical in the azide-treated Cp. The spectral properties of Cp provided by copper ions of different types change simultaneously and concordantly upon oxidation/reduction. This seems to be caused by cooperative interaction of these ions involved in the electron transfer from the donating substrate to the accepting molecule of the nitroxyl radical (in model studies of oxidase reaction) or oxygen (under natural conditions). The copper ions in the active centre of Cp constitute an intramolecular electron transport chain, which may, at least in vitro, function without one of its links.  相似文献   

15.
Summary Higher plant plasma membranes contain ab-type cytochrome that is rapidly reduced by ascorbic acid. The affinity towards ascorbate is 0.37 mM and is very similar to that of the chromaffin granule cytochromeb 561. High levels of cytochromeb reduction are reached when ascorbic acid is added either on the cytoplasmic or cell wall side of purified plasma membrane vesicles. This result points to a transmembrane organisation of the heme protein or alternatively indicates the presence of an effective ascorbate transport system. Plasma membrane vesicles loaded by ascorbic acid are capable of reducing extravesicular ferricyanide. Addition of ascorbate oxidase or washing of the vesicles does not eliminate this reaction, indicating the involvement of the intravesicular electron donor. Absorbance changes of the cytochromeb -band suggest the electron transfer is mediated by this redox component. Electron transport to ferricyanide also results in the generation of a membrane potential gradient as was demonstrated by using the charge-sensitive optical probe oxonol VI. Addition of ascorbate oxidase and ascorbate to the vesicles loaded with ascorbate results in the oxidation and subsequent re-reduction of the cytochromeb. It is therefore suggested that ascorbate free radical (AFR) could potentially act as an electron acceptor to the cytochrome-mediated electron transport reaction. A working model on the action of the cytochrome as an electron carrier between cytoplasmic and apoplastic ascorbate is discussed.Abbreviations AFR ascorbate free radical - AO ascorbate oxidase - DTT dithiothreitol - FCCP carbonylcyanidep-trifluorome-thoxyphenylhydrazon - Hepes N-(2-hydroxyethyl)-piperazine-N-(2-ethanesulfonic acid) - Oxonol VI bis(3-propyl-5-oxoisoxazol-4-yl) penthamethine oxonol - PMSF phenylmethylsulfluoride  相似文献   

16.
Both respiratory-competent and respiratory-deficient yeast cells reduce external ferricyanide. The reduction is stimulated by ethanol and inhibited by the alcohol dehydrogenase inhibitor, pyrazole. The reduction of ferricyanide is not inhibited by inhibitors of mitochondrial or microsomal ferricyanide reduction. Cells in exponential-phase growth show a much higher rate of ferricyanide reduction. The reduction of ferricyanide is accompanied by increased release of protons by the yeast cells. We propose that the ferricyanide reduction is carried out by a transmembrane NADH dehydrogenase.  相似文献   

17.
Some types of secretory vesicles, such as the chromaffin vesicles of the adrenal medulla, have cytochrome b561 which is believed to mediate the transfer of electrons across the vesicle membrane. To characterize the kinetics of this process, we have examined the rate of electron transfer from ascorbate trapped within chromaffin vesicle ghosts to external ferricyanide. The rate of ferricyanide reduction saturates at high ferricyanide concentrations. The reciprocal of the rate is linearly related to the reciprocal of the ferricyanide concentration. The internal ascorbate concentration affects the y intercept of this double-reciprocal plot but not the slope. These observations and theoretical considerations indicate that the slope is associated with a rate constant k1 for the oxidation of cytochrome b561 by ferricyanide. The intercept is associated with a rate constant k0 for the reduction of cytochrome b561 by internal ascorbate. From k0 and standard reduction potentials, the rate constant k-0 for the reduction of internal semidehydroascorbate by cytochrome b561 can be calculated. Under conditions prevailing in vivo, this rate of semidehydroascorbate reduction appears to be much faster than the expected rate of semidehydroascorbate disproportionation. This supports the hypothesis that cytochrome b561 functions in vivo to reduce intravesicular semidehydroascorbate thereby maintaining intravesicular ascorbic acid.  相似文献   

18.
Kipp BH  Kelley PM  Njus D 《Biochemistry》2001,40(13):3931-3937
Cytochrome b(561) mediates equilibration of the ascorbate/semidehydroascorbate redox couple across the membranes of secretory vesicles. The cytochrome is reduced by ascorbic acid and oxidized by semidehydroascorbate on either side of the membrane. Treatment with diethyl pyrocarbonate (DEPC) inhibits reduction of the cytochrome by ascorbate, but this activity can be restored by subsequent treatment with hydroxylamine, suggesting the involvement of an essential histidine residue. Moreover, DEPC inactivates cytochrome b(561) more rapidly at alkaline pH, consistent with modification of a histidine residue. DEPC does not affect the absorption spectrum of cytochrome b(561) nor does it change the midpoint reduction potential, confirming that histidine modification does not affect the heme. Ascorbate protects the cytochrome from inactivation by DEPC, indicating that the essential histidine is in the ascorbate-binding site. Further evidence for this is that DEPC treatment inhibits oxidation of the cytochrome by semidehydroascorbate but not by ferricyanide. This supports a reaction mechanism in which ascorbate loses a hydrogen atom by donating a proton to histidine and transferring an electron to the heme.  相似文献   

19.
Phytoplankton plasma membrane electron transport activity was determined by monitoring the reduction of the impermeant artificial electron acceptor ferricyanide in a range of diatoms. The results revealed that constitutive plasma membrane electron transport activity of marine diatoms is high compared with chlorophytes and higher plant cells. Diatom plasma membrane electron transport activity was not significantly increased by iron limitation. This lack of induction on iron limitation indicates that diatoms have an iron acquisition strategy that is distinct from chlorophytes and the dicotyledon higher plants that exhibit marked increases in plasma membrane ferricyanide reductase activity on iron limitation. The interaction of the constitutive plasma membrane electron transport with photosynthesis was also investigated. We found that 1) ferricyanide reduction at the plasma membrane was progressively inhibited in response to increasing irradiances; 2) the presence of extracellular ferricyanide, but not the reduced couple ferrocyanide, caused a marked inhibition of carbon fixation at high irradiance; and 3) extracellular electron acceptors ferricyanide and hexachloroiridate (but not ferrocyanide) induced an immediate and reversible decrease in fluorescence yields (Fo and Fm). The extent to which extracellular electron acceptors affected CO2 fixation, Fo, and Fm was related to the level of constitutive ferricyanide reductase activity, the species with highest ferricyanide reduction rates being most sensitive. The data suggest that consumption of electrons and/or reductant at the plasma membrane by external acceptors may compete directly with CO2 fixation for electrons, alter cytosolic‐chloroplast redox poise, and/or induce a redox‐signaling cascade that alters photosynthetic metabolism.  相似文献   

20.
The involvement of cytochrome b561, an integral membrane protein, in electron transfer across chromaffin-vesicle membranes is confirmed by changes in its redox state observed as changes in the absorption spectrum occurring during electron transfer. In ascorbate-loaded chromaffin-vesicle ghosts, cytochrome b561 is nearly completely reduced and exhibits an absorption maximum at 561 nm. When ferricyanide is added to a suspension of these ghosts, the cytochrome becomes oxidized as indicated by the disappearance of the 561 nm absorption. If a small amount of ferricyanide is added, it becomes completely reduced by electron transfer from intravesicular ascorbate. When this happens, cytochrome b561 returns to its reduced state. If an excess of ferricyanide is added, the intravesicular ascorbate becomes exhausted and the cytochrome b561 remains oxidized. The spectrum of these absorbance changes correlates with the difference spectrum (reduced-oxidized) of cytochrome b561. Cytochrome b561 becomes transiently oxidized when ascorbate oxidase is added to a suspension of ascorbate-loaded ghosts. Since dehydroascorbate does not oxidize cytochrome b561, it is likely that oxidation is caused by semidehydroascorbate generated by ascorbate oxidase acting on free ascorbate. This suggests that cytochrome b561 can reduce semidehydroascorbate and supports the hypothesis that the function of cytochrome b561 in vivo is to transfer electrons into chromaffin vesicles to reduce internal semidehydroascorbate to ascorbate.  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