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 359 毫秒
1.
Diferric transferrin which is often necessary for growth of cells is reduced by the transplasma membrane electron transport system of HeLa cells with release of ferrous iron outside the cell. Reduction of external diferric transferrin is reflected in oxidation of internal NADH. Adriamycin, an antitumor drug, inhibits diferric transferrin reduction by the HeLa cells and inhibits concomittant oxidation of cytosolic NADH at concentrations, 10(-8)-10(-6)M, which inhibit cell growth. Isolated liver plasma membranes have an NADH diferric transferrin reductase activity which is inhibited by similar adriamycin concentrations. We propose that inhibition of cell growth by adriamycin can be based on inhibition of transplasmalemma diferric transferrin reductase.  相似文献   

2.
Retinoic acid inhibits the reduction of diferric transferrin through the transplasma membrane electron transport system on fetal rat liver cells infected with a temperature-sensitive SV40 virus when the cells are in the nontransformed state cultured at 40°C. When the cells are in the transformed state (grown at the permissive 33°C temperature), retinoic acid does not inhibit the diferric transferrin reduction. Inhibition of activity of nontransformed cells is specific for retinoic acid with only slight inhibition by retinol and retinyl acetate at higher concentrations. Isolated rat liver plasma membrane NADH diferric transferrin reductase is also inhibited by retinoic acid. The effect of transformation with SV40 virus to decrease susceptibility to retinoic acid inhibition stands in contrast to much greater adriamycin inhibition of diferric transferrin reduction in the transformed cells than in nontransformed cells.  相似文献   

3.
Retinoic acid inhibition of transplasmalemma diferric transferrin reductase   总被引:1,自引:0,他引:1  
All trans retinoic acid inhibited diferric transferrin reduction by HeLa cells. The NADH diferric transferrin reductase activity of isolated liver plasma membranes was also inhibited by retinoic acid. Retinol and retinyl acetate had very little effect. Transplasma membrane ferricyanide reduction by HeLa cells and NADH ferricyanide reductase of liver plasma membrane was also inhibited by retinoic acid, therefore the inhibition was in the electron transport system and not at the transferrin receptor. Since the transmembrane electron transport has been shown to stimulate cell growth, the growth inhibition by retinoic acid thus may be based on inhibition of the NADH diferric transferrin reductase.  相似文献   

4.
Conjugates of adriamycin crosslinked to transferrin with glutaraldehyde inhibit proliferation of transformed cells. Conjugates of this type inhibit oxidoreductase activity in the plasma membrane of K562 cells, and the inhibition of electron transport is found at concentrations ten times lower than concentrations of free adriamycin which inhibit electron transport and cell growth. The transferrin-adriamycin conjugate inhibits ferricyanide reduction, diferric transferrin reduction and plasma membrane NADH oxidase activity stimulated by transferrin. Activation of proton release from the K562 cells by diferric transferrin also is inhibited by the conjugate, and conjugate kills cells more effectively than free adriamycin. Since the conjugate does not transfer adriamycin to the nucleus, the growth control may be based on inhibition of the transferrin regulated redox system and Na+/H+ antiport activity at the plasma membrane.  相似文献   

5.
NADH diferric transferrin reductase in liver plasma membrane   总被引:6,自引:0,他引:6  
Evidence is presented that rat liver plasma membranes contain a distinct NADH diferric transferrin reductase. Three different assay procedures for demonstration of the activity are described. The enzyme activity is highest in isolated plasma membrane, and activity in other internal membranes is one-eighth or less than in plasma membrane. The activity is inhibited by apotransferrin and antitransferrin antibodies. Trypsin treatment of the membranes leads to rapid loss of the transferrin reductase activity as compared with NADH ferricyanide reductase activity. Erythrocyte plasma membranes, which lack transferrin receptors, show no diferric transferrin reductase activity, although NADH ferricyanide reductase is present. The transferrin reductase is inhibited by agents that inhibit diferric transferrin reduction by intact cells and is activated by CHAPS (3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfate) detergent. Inhibitors of mitochondrial electron transport have no effect on the activity. We propose that the NADH diferric transferrin reductase in plasma membranes measures the activity of the enzyme that causes the reduction of diferric transferrin by intact cells. This transmembrane electron transport system requires the transferrin receptor for diferric transferrin reduction. Because the transmembrane electron transport has been shown to stimulate cell growth, the reduction of diferric transferrin at the cell surface may be an important function for diferric transferrin in stimulation of cell growth, in addition to its role in iron transport.  相似文献   

6.
Intact cells are known to reduce external, impermeable electron acceptors. We now show that cells can reduce the iron in diferric transferrin at the cell surface and that this reduction reaction depends on the transferrin receptor as well as the transmembrane electron transport system. Reduction of external diferric transferrin is accompanied by oxidation of internal NADH which indicates that the transmembrane enzyme is an NADH diferric transferrin reductase. Highly purified liver plasma membranes have NADH diferric transferrin reductase activity which shows properties similar to the diferric transferrin reductases activity of intact cells. Cell growth stimulation by diferric transferrin and other impermeable oxidants which can react with the diferric transferrin reductase can be based on electron transport through he plasma membrane.  相似文献   

7.
Proton release from HeLa cells is stimulated by external oxidants for the transplasmalemma electron transport enzymes. These oxidants, such as ferricyanide and diferric transferrin, also stimulate cell growth. We now present evidence that proton release associated with the reduction of ferricyanide and diferric transferrin is through the Na+/H+ antiport. The stoichiometry of H+/e- release with diferric transferrin is over 50 to 1, which is greater than expected for oxidation of a protonated transmembrane electron carrier. Diferric transferrin induced proton release depends on external sodium and is inhibited by amiloride. Proton release is also inhibited when diferric transferrin reduction is inhibited by apotransferrin. A tightly coupled association between the redox system and the antiport is shown by sodium dependence and amiloride inhibition of diferric transferrin reduction. The results indicate a new role for ferric transferrin in growth stimulation by activation of the sodium-proton antiport.  相似文献   

8.
Transplasmalemma electron transport by HeLa and pineal cells to reduce external ferricyanide is associated with proton release from the cells. Diferric transferrin also acts as an electron acceptor for the transmembrane oxidoreductase. We now show that reduction of external diferric transferrin by RPNA-209-1 SV40 transformed pineal cells is accompanied by proton release from the cells. The stoichiometry of proton release to electron transfer is much greater than would be expected from aniostropic electron flow across the membrane through protonated carriers. The proton release is not stimulated by apotransferrin and the diferric transferrin-stimulated activity is inhibited by apotransferrin. Apotransferrin also inhibits reduction of diferric transferrin by these cells. The proton release is dependent on external sodium ions and is inhibited by amiloride, which indicates that the proton release is mediated by the Na+/H+ antiport and that this antiport is activated by electron transport through the transmembrane dehydrogenase. Growth stimulation by diferric transferrin or other external oxidants can be based in part on activation of the Na+/H+ antiport.  相似文献   

9.
NADH oxidase of purified plasma membranes (electron transfer from NADH to oxygen) was stimulated by the growth factor diferric transferrin. This stimulation was of an activity not inhibited by cyanide and was not seen in plasma membranes prepared from hyperplastic nodules from liver of animals fed the hepatocarcinogen, 2-acetylaminofluorene, nor was it due to reduction of iron associated with diferric transferrin. With plasma membranes from nodules, the activity was already elevated and the added transferrin was without effect. The stimulation by diferric transferrin did not correlate with the absence of transferrin receptors which were increased at the nodule plasma membranes. With liver plasma membranes, the stimulation by diferric transferrin raised the plasma membrane NADH oxidase specific activity to approximately that of the nodule plasma membranes. In contrast to NADH oxidase, which was markedly stimulated by the diferric transferrin, NADH ferricyanide oxidoreductase or reduction of ferric ammonium citrate by liver plasma membranes was approximately equal to or slightly greater than that of the nodule plasma membrane and unaffected by diferric transferrin. The results suggest the possibility of coupling of NADH oxidase activity to a growth factor response in mammalian cells as observed previously for this enzyme in another system.  相似文献   

10.
Reduction of iron in diferric transferrin is inhibited by monoclonal antibodies to the transferrin receptor which bind at sites other than the high affinity transferrin binding site. These antibodies include B3/25, GB16 and GB22. Two antibodies which bind at the high affinity site for transferrin, 42/6 and GB18, do not inhibit iron reduction by transplasma membrane electron transport. The results are consistent with the proposal that differric transferrin reduction or stimulation of transmembrane NADH oxidase activity involves a site different from the high affinity diferric transferrin binding site. A synergistic action of antibodies with epitopes at the tight binding site involved in iron uptake and the antibodies which inhibit electron transport, B3/25 and GB16, can explain the increased inhibition of growth observed when both 42/6 and B3/25 are added to proliferating cells.  相似文献   

11.
The transfer of iron from diferric transferrin to bathophenanthroline disulfonate was measured under varying conditions by spectrophotometry and EPR spectroscopy. Intact rat hepatocytes efficiently mediated the transfer of iron from human diferric transferrin to bathophenanthroline disulfonate. Isolated rat liver plasma membranes, in contrast, failed to facilitate the reaction at pH 7.4 in the presence of NADH, although the membranes were able to reduce ferricyanide and to oxidize NADH. Oxidation of NADH was stimulated by diferric transferrin. However, ferricyanide reductase and transferrin-stimulated NADH oxidase activities were apparently not linked to release of iron from transferrin. Our results, together with theoretical considerations, show that the ability (or inability) of intact cells or isolated plasma membranes to facilitate the transfer of iron from transferrin to strong diferric iron chelators does not allow interferences about the existence of an iron reduction step as part of the process of cellular uptake of iron from transferrin.  相似文献   

12.
NADH oxidase activity (electron transfer from NADH to molecular oxygen) of plasma membranes purified from rat liver was characterized by a cyanide-insensitive rate of 1 to 5 nmol/min per mg protein. The activity was stimulated by growth factors (diferric transferrin and epidermal growth factor) and hormones (insulin and pituitary extract) 2- to 3-fold. In contrast, NADH oxidase was inhibited up to 80% by several agents known to inhibit growth or induce differentiation (retinoic acid, calcitriol, and the monosialoganglioside, GM3). The growth factor-responsive NADH oxidase of isolated plasma membranes was not inhibited by common inhibitors of oxidoreductases of endoplasmic reticulum or mitochondria. As well, NADH oxidase of the plasma membrane was stimulated by concentrations of detergents which strongly inhibited mitochondrial NADH oxidases and by lysolipids or fatty acids. Growth factor-responsive NADH oxidase, however, was inhibited greater than 90% by chloroquine and quinone analogues. Addition of coenzyme Q10 stimulated the activity and partially reversed the analogue inhibition. The pH optimum for NADH oxidase was 7.0 both in the absence and presence of growth factors. The Km for NADH was 5 microM and was increased in the presence of growth factors. The stoichiometry of the electron transfer reaction from NADH to oxygen was 2 to 1, indicating a 2 electron transfer. NADH oxidase was separated from NADH-ferricyanide reductase, also present at the plasma membrane, by ion exchange chromatography. Taken together, the evidence suggests that NADH oxidase of the plasma membrane is a unique oxidoreductase and may be important to the regulation of cell growth.  相似文献   

13.
Transplasma membrane electron transport from HeLa cells, measured by reduction of ferricyanide or diferric transferrin in the presence of bathophenanthroline disulfonate, is inhibited by low concentrations of adriamycin and adriamycin conjugated to diferric transferrin. Inhibition with the conjugate is observed at one-tenth the concentration required for adriamycin inhibition. The inhibitory action of the conjugate appears to be at the plasma membrane since (a) the conjugate does not transfer adriamycin to the nucleus, (b) the inhibition is observed within three minutes of addition to cells, and (c) the inhibition is observed with NADH dehydrogenase and oxidase activities of isolated plasma membranes. Cytostatic effects of the compounds on HeLa cells show the same concentration dependence as for enzyme inhibition. The adriamycin-ferric transferrin conjugate provides a more effective tool for inhibition of the plasma membrane electron transport than is given by the free drug.  相似文献   

14.
The short-term incubation of HeLa cells in the presence of diferric transferrin or ferricyanide, which are reduced externally by the transplasma membrane reductase, produces a stoichiometric decrease in NADH and increase in NAD+, which is stimulated by insulin. The NADP/NADPH ratio does not change during 15 min incubation with the oxidants. The total pyridine nucleotide pool of HeLa cells is not affected. Incubation with apotransferrin and ferrocyanide, which cannot act as oxidants for transmembrane electron transport, does not change the pyridine nucleotide concentrations in the cells. Our results show that NADH can act as the internal electron donor for the reduction of external oxidants by the transmembrane reductase. It appears that oxidation of NADH by the transmembrane electron transport using ferricyanide or iron transferrin as external electron acceptors is sufficient to stimulate growth in HeLa cells.  相似文献   

15.
The presence of yeast cells in the incubation medium prevents the oxidation of ascorbate catalyzed by copper ions. Ethanol increases ascorbate retention. Pyrazole, an alcohol dehydrogenase inhibitor, prevents ascorbate stabilization by cells. Chelation of copper ions does not account for stabilization, since oxidation rates with broken or boiled cells or conditioned media are similar to control rates in the absence of cells. Protoplast integrity is needed to reach optimal values of stabilization. Chloroquine, a known inhibitor of plasma membrane redox systems, inhibits the ascorbate stabilization, the inhibition being partially reversed by coenzyme Q6. Chloroquine does not inhibit ferricyanide reduction. Growth of yeast in iron-deficient media to increase ferric ion reductase activity also increases the stabilization. In conclusion, extracellular ascorbate stabilization by yeast cells can reflect a coenzyme Q dependent transplasmalemma electron transfer which uses NADH as electron donor. Iron deficiency increases the ascorbate stabilization but the transmembrane ferricyanide reduction system can act independently of ascorbate stabilization.  相似文献   

16.
Rat liver ferritin is an effective donor of iron to rat hepatocytes. Uptake of iron from ferritin by the cells is partially inhibited by including apotransferrin in the culture medium, but not by inclusion of diferric transferrin. This inhibition is dependent on the concentration of apotransferrin, with a 30% depression in iron incorporation in the cells detected at apotransferrin concentrations above 40 micrograms/ml. However, apotransferrin does not interfere with uptake of 125I-labeled ferritin, suggesting that apotransferrin decreases retention of iron taken up from ferritin by hepatocytes by sequestering a portion of released iron before it has entered the metabolic pathway of the cells. The iron chelators desferrioxamine (100 microM), citrate (10 mM) and diethylenetriaminepentaacetate (100 microM) reduce iron uptake by the cells by 35, 25 and 8%, respectively. In contrast, 1 mM ascorbate increases iron accumulation by 20%. At a subtoxic concentration of 100 microM, chloroquine depresses ferritin and iron uptake by hepatocytes by more than 50% after 3 h incubation. Chloroquine presumably acts by retarding lysosomal degradation of ferritin and recycling of ferritin receptors.  相似文献   

17.
Chloroquine inhibits the incorporation of [14C]acetate into sterols at a concentration of 10 microM or more in mouse L cells but has no effect on fatty acid synthesis and CO2 production from the same substrate even at a 10-fold higher concentration of the drug. The site of inhibition is distal to the formation of mevalonate since chloroquine also inhibits [14C]mevalonate metabolism to sterols and does not decrease the activity of 3-hydroxy-3-methylglutaryl coenzyme A reductase (EC 1.1.1.34) or the incorporation of [14C]acetate into the total nonsaponifiable lipids. Analyses by thin layer and high pressure liquid chromatography of the nonsaponifiable lipid fraction from cultures incubated with chloroquine show an accumulation of radioactivity in the region of squalene oxide. Identification of the radiolabeled lipid as squalene oxide has been established by: (a) its co-migration with the authentic squalene oxide standard; (b) its conversion into squalene glycol by acid hydrolysis; and (c) its further metabolism to desmosterol when chloroquine is removed from the medium. Addition of chloroquine (12.5-50 microM) to 20,000 X g supernatant fractions of mouse liver homogenates inhibits the incorporation of [14C]mevalonolactone into cholesterol and lanosterol, with corresponding increases of [14C]squalene oxides, in a concentration-dependent manner. It appears, therefore, that chloroquine inhibits the enzymatic step catalyzed by 2,3-oxidosqualene-lanosterol cyclase (EC 5.4.99.7). Incubation of cell cultures with chloroquine (50 microM) arrests cell growth and causes cell death after 1-3 days. However, simultaneous incubation of chloroquine with either cholesterol or lanosterol prevents cell death and permits cell growth. Uptake of chloroquine is not affected by exogenous sterols since intracellular chloroquine concentrations are the same in cells grown with or without added sterols. The cytotoxicity of chloroquine, under our experimental conditions, must, therefore, be due primarily to its inhibition of sterol synthesis. In addition to its well known effect on protein catabolism, chloroquine has been found to inhibit protein synthesis. The significance of these findings concerning the use of chloroquine in studying the regulation of 3-hydroxy-3-methylglutaryl coenzyme A reductase activity is discussed.  相似文献   

18.
The reductant dependence of iron mobilization from isolated rabbit reticulocyte endosomes containing diferric transferrin is reported. The kinetic effects of acidification by a H(+)-ATPase are eliminated by incubating the endosomes at pH 6.0 in the presence of 15 microM FCCP to acidify the intravesicular milieu and to dissociate 59Fe(III) from transferrin. In the absence of reductants, iron is not released from the vesicles, and iron leakage is negligible. The second-order dependence of rate constants and amounts of 59Fe mobilized from endosomes using ascorbate, ferrocyanide, or NADH are consistent with reversible mechanisms. The estimated apparent first-order rate constant for mobilization by ascorbate is (2.7 +/- 0.4) x 10(-3) s-1 in contrast to (3.2 +/- 0.1) x 10(-4) s-1 for NADH and (3.5 +/- 0.6) x 10(-4) s-1 for ferrocyanide. These results support models where multiple reactions are involved in complex processes leading to iron transfer and membrane translocation. A type II NADH dehydrogenase (diaphorase) is present on the endosome outer membrane. The kinetics of extravesicular ferricyanide reduction indicate a bimolecular-bimolecular steady-state mechanism with substrate inhibition. Ferricyanide inhibition of 59Fe mobilization is not detected. Significant differences between mobilization and ferricyanide reduction kinetics indicate that the diaphorase is not involved in 59Fe(III) reduction. Sequential additions of NADH followed by ascorbate or vice versa indicate a minimum of two sites of 59Fe(III) residence; one site available to reducing equivalents from ascorbate and a different site available to NADH. Sequential additions using ferrocyanide and the other reductants suggest interactions among sites available for reduction. Inhibition of ascorbate-mediated mobilization by DCCD and enhancement of ferrocyanide and NADH-mediated mobilization suggest a role for a moiety with characteristics of a proton pore similar to that of the H(+)-ATPase. These data provide significant constraints on models of iron reduction, translocation, and mobilization by endocytic vesicles.  相似文献   

19.
Transformation of 3T3 cells by SV40 virus changes the properties of the transplasma membrane electron transport activity which can be assayed by reduction of external ferric salts. After 42 h of culture and before the growth rate is maximum, the transformed cells have a much slower rate of ferric reduction. The change in activity is expressed both by change inK m andV max for ferricyanide reduction. The change in activity is not based on surface charge effect or on tight coupling to proton release or on intracellular NADH concentration. With transformation by SV40 virus infection the expression of transferrin receptors increases, which correlates with greater diferric transferrin stimulation of the rate of ferric ammonium citrate reduction in transformed SV40-3T3 cells than in 3T3 cells.  相似文献   

20.
Summary Chloroquine is an antimalarial and antirheumatic lysosomotropic drug which inhibits taurine uptake into and increases efflux from cultured human lymphoblastoid cells. It inhibits taurine uptake by rat lung slices and affects the uptake and release of cystine from cystinotic fibroblasts. Speculations on its mode of action include a proton gradient effect, a non-specific alteration in membrane integrity, and membrane stabilization. In this study, the effect of chloroquine on the uptake of several amino acids by rat renal brush border membrane vesicles (BBMV) was examined. Chloroquine significantly inhibited the secondary active, NaCl-dependent component of 10µM taurine uptake at all concentrations tested, but did not change equilibrium values. Analysis of these data indicated that the inhibition was non-competitive. Taurine uptake was reduced at all osmolarities tested, but inhibition was greatest at the lowest osmolarity. Taurine efflux was not affected by chloroquine, nor was the NaCl-independent diffusional component of taurine transport. Chloroquine (1 mM) inhibited uptake of the imino acids L-proline and glycine, and the dibasic amino acid L-lysine. It inhibited the uptake of D-glucose, but not the neutral-amino acids L-alanine or L-methionine. Uptake of the dicarboxylic amino acids, L-glutamic acid and L-aspartic acid, was slightly enhanced. With regard to amino acid uptake by BBMV, these findings may support some of the currently proposed mechanisms of the action of chloroquine but further studies are indicated to determine why it affects the initial rate of active amino acid transport.  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