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.
The addition of inositol to the growth medium of Saccharomyces cerevisiae resulted in rapid changes in the rates of phospholipid biosynthesis. The partitioning of the phospholipid intermediate CDP-diacylglycerol was shifted to phosphatidylinositol at the expense of phosphatidylserine and its derivatives phosphatidylethanolamine and phosphatidylcholine. Serine at 133-fold greater concentrations than that of inositol shifted the partitioning of CDP-diacylglycerol to phosphatidylserine at the expense of phosphatidylinositol but to a much lesser degree. Kinetic experiments with pure phosphatidylserine synthase and phosphatidylinositol synthase indicated that the partitioning of CDP-diacylglycerol between phosphatidylserine and phosphatidylinositol was not governed by the affinities both enzymes have for their common substrate CDP-diacylglycerol. Instead, the main regulation of phosphatidylinositol and phosphatidylserine synthesis was through the exogenous supply of inositol. The Km of inositol (0.21 mM) for phosphatidylinositol synthase was 9-fold higher than cytosolic concentration of inositol (24 microM). The Km of serine (0.83 mM) for phosphatidylserine synthase was 3-fold below the cytosolic concentration of serine (2.6 mM). Therefore, inositol supplementation resulted in a dramatic increase in the rate of phosphatidylinositol synthesis, whereas serine supplementation resulted in little affect on the rate of phosphatidylserine synthesis. Inositol also contributed to the regulation of phosphatidylinositol and phosphatidylserine synthesis by having a direct affect on phosphatidylserine synthase activity. Kinetic experiments with pure phosphatidylserine synthase showed that inositol was a noncompetitive inhibitor of the enzyme with a Ki of 65 microM.  相似文献   

2.
The in vitro effect of bovine brain cortex phosphatidylserine on 32Pi incorporation into phosphatidylinositol, phosphatidylcholine, and phosphatidylethanolamine of rat anterior pituitary glands was studied. Phosphatidylserine (0.1 to 66.6 microM) decreased the incorporation of 32Pi into phosphatidylinositol, but not phosphatidylcholine or phosphatidylethanolamine, in a concentration-related manner. The inhibitory effect of phosphatidylinositol was similar to that of dopamine in the same experimental conditions. The combined effects of submaximal concentrations of dopamine and phosphatidylserine elicited an apparently additive inhibitory effect on phosphatidylinositol synthesis. The inhibitory effect of phosphatidylserine was completely reversed by haloperidol and sulpiride and only partially by pimozide, antidopaminergic agents which per se do not affect phosphatidylinositol synthesis. The stimulatory effect of TRH to increase 32Pi incorporation into phosphatidylinositol was decreased by phosphatidylserine. These observations suggest that the decrease in prolactin release in the presence of phosphatidylserine may be evoked through a dopaminergic mechanism.  相似文献   

3.
The effect of phosphatidylserine exogenously added to the medium on de novo biosynthesis of phosphatidylserine was investigated in cultured Chinese hamster ovary cells. When cells were cultured for several generations in medium supplemented with phosphatidylserine and 32Pi, the incorporation of 32Pi into cellular phosphatidylserine was remarkably inhibited, the degree of inhibition being dependent upon the concentration of added phosphatidylserine. 32Pi uptake into cellular phosphatidylethanolamine was also partly reduced by the addition of exogenous phosphatidylserine, consistent with the idea that phosphatidylethanolamine is biosynthesized via decarboxylation of phosphatidylserine. However, incorporation of 32Pi into phosphatidylcholine, sphingomyelin, and phosphatidylinositol was not significantly affected. In contrast, the addition of either phosphatidylcholine, sphingomyelin, phosphatidylethanolamine, or phosphatidylinositol to the medium did not inhibit endogenous biosynthesis of the corresponding phospholipid. Radiochemical and chemical analyses of the cellular phospholipid composition revealed that phosphatidylserine in cells grown with 80 microM phosphatidylserine was almost entirely derived from the added phospholipid. Phosphatidylserine uptake was also directly determined by using [3H]serine-labeled phospholipid. Pulse and pulse-chase experiments with L-[U-14C] serine showed that when cells were cultured with 80 microM phosphatidylserine, the rate of synthesis of phosphatidylserine was reduced 3-5-fold whereas the turnover of newly synthesized phosphatidylserine was normal. Enzyme assaying of extracts prepared from cells grown with and without phosphatidylserine indicated that the inhibition of de novo phosphatidylserine biosynthesis by the added phosphatidylserine appeared not to be caused by a reduction in the level of the enzyme involved in the base-exchange reaction between phospholipids and serine. These results demonstrate that exogenous phosphatidylserine can be efficiently incorporated into Chinese hamster ovary cells and utilized for membrane biogenesis, endogenous phosphatidylserine biosynthesis thereby being suppressed.  相似文献   

4.
Ca2+-induced phase separation in phosphatidylserine/phosphatidylethanolamine and phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine model membranes was studied using spin-labeled phosphatidylethanolamine and phosphatidylcholine and compared with that in phosphatidylserine/phosphatidylcholine model membranes studied previously. The phosphatidylethanolamine-containing membranes behaved in qualitatively the same way as did phosphatidylserine/phosphatidylcholine model membranes. There were some quantitative differences between them. The degree of phase separation was higher in the phosphatidylethanolamine-containing membranes. For example, the degree of phase separation in phosphatidylserine/phosphatidylethanolamine membranes containing various mole fractions of phosphatidylserine was 94--100% at 23 degrees C and 84--88% at 40 degrees C, while the corresponding value for phosphatidylserine/phosphatidylcholine membranes was 74--85% at 23 degrees C and 61--79% at 40 degrees C. Ca2+ concentration required for the phase separation was lower for phosphatidylserine/phosphatidylethanolamine than that for phosphatidylserine/phosphatidylcholine membranes; concentration to cause a half-maximal phase separation was 1.4 . 10(-7) M for phosphatidylserine-phosphatidylethanolamine and 1.2 . 10(-6) M for phosphatidylserine/phosphatidylcholine membranes. The phase diagram of phosphatidylserine/phosphatidylethanolamine membranes in the presence of Ca2+ was also qualitatively the same as that of phosphatidylserine/phosphatidylcholine except for the different phase transition temperatures of phosphatidylethanolamine (17 degrees C) and phosphatidylcholine (-15 degrees C). These differences were explained in terms of a greater tendency for phosphatidylethanolamine, compared to phosphatidylcholine, to form its own fluid phase separated from the Ca2+-chelated solid-phase phosphatidylserine domain.  相似文献   

5.
The effect of norepinephrine on phosphatidylcholine and phosphatidylethanolamine formation was investigated in short-term incubations with freshly isolated rat hepatocytes. In the presence of dl-propranolol, norepinephrine decreases the incorporation of [methyl-14C]choline into phosphatidylcholines in a dose-dependent manner. At a concentration of 50 microM, norepinephrine (plus 20 microM propranolol) inhibits the incorporation of [methyl-14C]choline over a wide range of choline concentrations (59% inhibition at 5 microM choline; 34% inhibition at 1 mM choline). Norepinephrine also decreases the incorporation rates of [1-14C]palmitic acid and [1-14C]oleic acid into phosphatidylcholines. The effect of norepinephrine is mediated through an alpha-adrenergic receptor. Norepinephrine (plus propranolol) does not decrease the uptake or phosphorylation rate of [methyl-14C]choline. Pulse-label and pulse-chase studies indicate that the conversion rate of phosphocholine to CDP-choline, catalyzed by CTP:phosphocholine cytidylyltransferase, is diminished by norepinephrine. In contrast with the inhibitory effect of norepinephrine on phosphatidylcholine synthesis, this hormone stimulates the formation of phosphatidylethanolamines from [1,2-14C]ethanolamine. This increased incorporation rate is apparent at ethanolamine concentrations above 25 microM. A combination of norepinephrine and propranolol decreases, however, the synthesis of phosphatidylcholines from [1,2-14C]ethanolamine. The results indicate that alpha-adrenergic regulation dissociates the synthesis of phosphatidylcholines from that of phosphatidylethanolamines.  相似文献   

6.
The Ca2(+)-ATPase of skeletal sarcoplasmic reticulum was purified and reconstituted in the presence of phosphatidyl choline using the freeze-thaw sonication technique. The effect of incorporation of negatively charged phospholipids, phosphatidylserine and phosphatidylinositol phosphate, into the phosphatidylcholine proteoliposomes was investigated. Various ratios of phosphatidylserine or phosphatidylinositol phosphate to phosphatidylcholine were used, while the total amount of phospholipid in the reconstituted vesicles was kept constant. Enrichment of phosphatidylcholine proteoliposomes by phosphatidylserine or phosphatidylinositol phosphate was associated with activation of Ca2(+)-uptake and Ca2(+)-ATPase activities. The highest activation was obtained at a 50:50 molar ratio of phosphatidylserine:phosphatidylcholine and at a 10:90 molar ratio of phosphatidylinositol phosphate:phosphatidylcholine. The initial rates of Ca2(+)-uptake obtained at 1 microM Ca2+ were 2.6 +/- 0.1 mumol/min per mg of phosphatidylserine:phosphatidylcholine proteoliposomes and 1.5 +/- 0.1 mumol/min per mg of phosphatidylinositol phosphate:phosphatidylcholine proteoliposomes, compared to 0.9 +/- 0.05 mumol/min per mg of phosphatidylcholine proteoliposomes. These findings suggest that negatively charged phospholipids may be involved in the activation of the reconstituted skeletal muscle sarcoplasmic reticulum Ca2(+)-pump.  相似文献   

7.
Propranolol, a beta-adrenergic receptor antagonist, also inhibits phosphatidate phosphohydrolase, the enzyme that converts phosphatidic acid into diacylglycerol. This latter effect has prompted recent use of propranolol in studies examining the importance of diacylglycerol and phosphatidic acid in cellular signalling events. Here, we show that propranolol is also an inhibitor of protein kinase C. At concentrations greater than or equal to 20 microM, propranolol reduced [3H]phorbol dibutyrate binding (IC50 = 200 microM) and phorbol myristate acetate-stimulated superoxide anion release (IC50 = 130 microM) in human neutrophils. Scatchard analysis showed that propranolol lowers the number of phorbol diester binding sites without significantly affecting their affinity. In vitro kinetic analysis, performed in a mixed micellar assay with protein kinase C purified from human neutrophils, suggested a competitive inhibition of propranolol with the cofactor phosphatidylserine. Complex kinetic patterns were observed with respect to diacylglycerol and ATP, approximating competitive and noncompetitive inhibition, respectively. Taken together, these results suggest that the drug interacts at the level of the regulatory domain of the enzyme. Fifty % inhibition occurred at approximately 150 microM propranolol. Similar levels of inhibition were obtained using exogenous (histone) and endogenous (p47-phox, a NADPH oxidase component) substrates. Protein kinase C-alpha and protein kinase C-beta, two protein kinase C isozymes present in human neutrophils, were inhibited by propranolol in a comparable manner. In the range of concentrations tested (30-1000 microM), neither cAMP-dependent protein kinase nor neutrophil protein tyrosine kinases were affected. The racemic form of propranolol and the (+) and the (-) stereoisomers were equally active, and other beta-adrenergic receptor antagonists (pindolol) and agonists (isoproterenol) were inactive. This suggests that the inhibitory action of propranolol on protein kinase C is related to the amphipathic nature of the drug rather than to its beta-adrenergic receptor blocking ability. Analogs of propranolol were synthesized and found to be more potent protein kinase C inhibitors, with IC50 values in the 10-20 microM range. We conclude that the ability of propranolol to inhibit both protein kinase C and PA phosphohydrolase complicates interpretation of results when this drug is used in signal transduction studies. In addition, propranolol may be a useful prototype for the synthesis of new protein kinase C inhibitors.  相似文献   

8.
Covalent linkage of ethylenediamine with the Na+/K(+)-ATPase complex from rabbit kidney outer medulla by the use of the water-soluble carbodiimide, N-ethyl,N'-(3-dimethylaminopropyl)carbodiimide, resulted in a 73% reaction with phosphatidylserine and only 27% with carboxylic groups in the proteic component of the enzyme. Condensation products from the reaction between phosphatidylserine and ethylenediamine, N-(O-phosphatidylseryl)ethylenediamine, N,N'-bis(O-phosphatidylseryl)ethylenediamine and its intermediary product O-phosphatidyl-[N,N'-bis(seryl)]ethylenediamine, were synthesised. Symmetrically substituted ethylenediamine was the most likely condensation product of ethylenediamine with endogenous phosphatidylserine. The synthesised lipids were incorporated in proteoliposomes containing Na+/K(+)-ATPase and only the addition of the phospholipid phosphatidylcholine. The ratio of phospholipid to protein was 52 (w/w). These proteoliposomes were perforated by the addition of 0.5% cholate and both the Na(+)-dependent phosphorylation level and its dependence on Na+, Mg2+ and ATP were measured. Phosphatidylcholine alone increased the half-maximal activation concentration for Na+ ([Na+]0.5) from 0.2 to 1-2 mM, for Mg2+ from 0.1 to 0.8 microM and for ATP from 0.02 to 0.3 microM. The Ki for K+ (in the absence of Na+) was unaffected: 12.8 microM vs. 12.5 microM in the non-reconstituted system. Replacing 10 mol% of phosphatidylcholine by phosphatidylethanolamine: or phosphatidylserine had no significant effect on [Na+]0.5: 1.1 and 0.7 mM, respectively. Replacing 5 mol% phosphatidylcholine by the bis(phosphatidylseryl) substituent of ethylenediamine further increased [Na+]0.5 to 13.7 mM, while half-maximal activation concentrations for Mg2+ and ATP were unaltered. The mono-phosphatidylseryl derivatives of ethylenediamine, each 5 mol%, also increased [Na+]0.5, but to a lesser extent (3.2-3.8 mM). In addition to their competitive effects, the phosphatidylseryl-substituted ethylenediamine compounds exerted a slowly-increasing non-competitive inhibition, not only in phosphorylation, but also in overall ATPase activity, which was reduced, although not abolished, by exogenous protein (bovine serum albumin). A detergent-like action in the usual sense is unlikely since liposomes containing these lipids remained intact. These studies prove that phospholipids are not only required for optimal activity of this transport enzyme, but in excess or in compositions deviating from the normal, may also be inhibitory.  相似文献   

9.
Incubation of Jurkat cells in the presence of H2O2 either directly added to the culture medium or generated with glucose oxidase, menadione or the couple xanthine/xanthine oxidase induced a marked decrease of phosphatidylserine synthesis in the absence of changes in the synthesis of phosphatidylcholine and phosphatidylethanolamine. Concentration dependent response curves indicated that H2O2 induced inhibition of phosphatidylserine synthesis with an IC(50)=5 microM while both induction of tyrosine phosphorylation of proteins and Ca(2+) signals were obtained with an EC(50)=300 microM. The tyrosine kinase and Ca(2+) independent mechanism was confirmed by comparing the H2O2-induced and the CD3-induced inhibition of phosphatidylserine synthesis using several Jurkat clones differing in the expression of cell surface receptors such as CD3/TCR and CD45 and protein tyrosine kinase such as p72syk, ZAP-70 and p56lck. While CD3-induced inhibition of phosphatidylserine synthesis necessitates protein tyrosine phosphorylation and Ca(2+) signals, H2O2 provoked its effect in all the clones studied independently of the presence or absence of the proteins previously shown to be key elements in T cell signal transduction. Conversely, the antioxidant molecule, butylated hydroxanisole, generates an increased PtdSer synthesis, suggesting that the synthesis of this phospholipid is regulated by the redox status of the cells.  相似文献   

10.
The bee and cobra venom phospholipases A2 as well as partially acetylated cobra venom phospholipase A2 are studied for their effect on phospholipid composition of synaptosomes and their Mg2+- and Na+,K+-ATPase activity. It is established that these phospholipases induce the splitting of phosphatidylethanolamine, phosphatidylcholine and phosphatidylserine, inhibition of the Na+,K+-ATPase activity and activation of Mg2+-ATPase. Bee venom phospholipase A2 is more effective than cobra venom phospholipase A2, the both phospholipases splitting phosphatidylethanolamine most intensively. The ATPase activity may be partially or completely restored by exogenic phosphatidylcholine and phosphatidylserine; exogenic phosphatidylethanolamine is not efficient in this respect.  相似文献   

11.
The effect of cationic amphiphilic drugs, chlorpromazine and dibucaine, on phospholipid metabolism in the frog sartorius muscle was studied at concentrations affecting excitation-contraction coupling. Labelling patterns of phospholipids using 32P orthophosphate were determined. The drugs at a concentration of 3.10(-4) mol/l inhibited the synthesis of phosphatidylcholine and phosphatidylethanolamine and stimulated the synthesis of phosphatidylinositol and phosphatidylserine. At the 3.10(-3) mol/l concentration the drugs blocked the synthesis of all phospholipids without appreciably affecting their degradation. The effect of cationic amphiphilic drugs was independent of the presence of Ca2+ or K+ ions in the media, however, basal labelling of phospholipids was affected by withdrawal of Ca2+ or K+ ions.  相似文献   

12.
On the substrate specificity of rat liver phospholipase A1   总被引:1,自引:0,他引:1  
The substrate specificity of purified phospholipase A1 was studied using mixed micelles of phospholipid and Triton X-100. The kinetic analysis employed determined Vmax, Ks (a dissociation constant for the phospholipase A1-mixed micelle complex), and Km (the Michaelis constant for the catalytic step which reflects the binding of the enzyme to the substrate in the interface). The order of Vmax values was phosphatidic acid greater than phosphatidylethanolamine greater than phosphatidylcholine greater than phosphatidylserine. The order of Ks values was phosphatidylcholine greater than phosphatidylethanolamine greater than phosphatidic acid greater than phosphatidylserine; the order of Km values was phosphatidic acid greater than phosphatidylethanolamine = phosphatidylserine greater than phosphatidylcholine. When present together, phosphatidylcholine inhibited the hydrolysis of phosphatidylethanolamine but phosphatidylethanolamine did not affect the hydrolysis of phosphatidylcholine. Sphingomyelin, phosphatidylcholine plasmalogen, and phosphatidylethanolamine plasmalogen had no effect on the hydrolysis of phosphatidylethanolamine. The effects of the reaction products, lysolipids and/or fatty acids, were also considered for their influence on phosphatidylethanolamine hydrolysis catalyzed by phospholipase A1. Free fatty acid was found to inhibit, whereas lysophospholipids stimulated hydrolysis of phosphatidylethanolamine. In a mixture of 1,2- and 1,3-diacylglycerides in mixed micelles, only the acyl chain at the sn-1 position of the 1,2 compound was hydrolyzed. Surface charge did not modulate the hydrolysis of phosphatidylcholine vesicles or mixed micelles. In conclusion, it is hypothesized that steric hindrance at position 3 of the glycerol regulates substrate binding in the active site and that an acyl group in position 1 is favored over a vinyl ether linkage for binding.  相似文献   

13.
Triacsins A, B, C, and D are new inhibitors of long chain acyl-CoA synthetase (EC 6.2.1.3) and possess different inhibitory potencies against the enzyme (Tomoda, H., Igarashi, K., and Omura, S. (1987) Biochim. Biophys. Acta 921, 595-598). Acyl-CoA synthetase activity in the membrane fraction of Raji cells was also inhibited by triacsins. The same hierarchy of inhibitory potency as that against the enzyme from other sources, triacsin C greater than triacsin A much greater than triacsin D greater than or equal to triacsin B, was observed. When Raji cells were cultivated in the presence of triacsins, cell proliferation was inhibited in a dose-dependent fashion. The drug concentrations required for 50% inhibition of cell growth at day 2 were calculated to be 1.8 microM for triacsin A, much greater than 20 microM for triacsin B, 1.0 microM for triacsin C, and much greater than 15 microM for triacsin D, demonstrating a hierarchy for inhibitory potency of triacsins similar to that against the acyl-CoA synthetase activity. To understand the role of long chain acyl-CoA synthetase in animal cells, the effect of triacsins on the lipid metabolism of Raji cells was studied. When intact Raji cells were incubated with [14C]oleate in the presence of individual triacsins, the incorporation of [14C]oleate into each of the lipid fractions such as phosphatidylcholine, phosphatidylethanolamine, and triacylglycerol was inhibited to an analogous extent. A common hierarchy, triacsin C greater than triacsin A much greater than triacsin D greater than triacsin B, was shown for the inhibition in each synthesis of the three lipids, which was identical with that for acyl-CoA synthetase. These findings indicate that the inhibition of acyl-CoA synthetase is well correlated with the inhibition of lipid synthesis. Taken together, the data strongly suggest that the inhibition of acyl-CoA synthetase by triacsins leads to the inhibition of lipid synthesis and eventually to the inhibition of proliferation of Raji cells.  相似文献   

14.
The "membrane bilayer" pathway (Rhodes, D. G., J. G. Sarmiento, and L. G. Herbette. 1985. Mol. Pharmacol. 27:612-623.) for 1,4-dihydropyridine calcium channel drug (DHP) binding to receptor sites in cardiac sarcolemmal membranes has been extended to include the interaction of amphiphiles within the lipid bilayer. These studies focused on the ability of the Class III antiarrhythmic agents bretylium and clofilium to nonspecifically inhibit DHP-receptor binding in canine cardiac sarcolemma. Clofilium was found to inhibit nimodipine binding with an inhibition constant of approximately 5 microM, whereas bretylium had no effect on nimodipine binding. Small angle x-ray diffraction was then used to examine the differential ability of these two Class III agents to inhibit DHP-receptor binding. The time-averaged locations of bretylium, clofilium, and nimodipine in bovine cardiac phosphatidylcholine (BCPC) bilayers (supplemented with 13 mol% cholesterol) were determined to a resolution of 9 A. The location of bretylium as dominated by its phenyl ring in BCPC bilayers was found to be at the hydrocarbon core/water interface, similar to that of the dihydropyridine ring of nimodipine. The location of clofilium as dominated by its phenyl ring was found to be below the hydrocarbon/core water interface within the hydrocarbon chain region of the bilayer, similar to that of the phenyl ring of nimodipine. The location of the dihydropyridine ring portion of nimodipine has previously been shown by neutron diffraction to be located at the hydrocarbon core/water interface of native sarcoplasmic reticulum, consistent with the small angle x-ray data from model membranes in this paper. Therefore, we speculate that the nonspecific inhibition arises from the interaction of clofilium's phenyl ring with the site on the calcium channel receptor where the phenyl ring portion of nimodipine must interact. The DHP-receptor binding pathway would then involve both nonspecific (membrane) and specific (protein) binding components, both of which are necessary for receptor binding.  相似文献   

15.
Ca2+-induced phase separation in phosphatidylserine/phosphatidylethanolamine and phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine model membranes was studied using spin-labeled phosphatidylethanolamine and phosphatidylcholine and compared with that in phosphatidylserine/phosphatidylcholine model membranes studied previously. The phosphatidyl-ethanolamine-containing membranes behaved in qualitatively the same way as did phosphatidylserine/phosphatidylcholine model membranes. There were some quantitative differences between them. The degree of phase separation was higher in the phosphatidylethanolamine-containing membranes. For example, the degree of phase separation in phosphatidylserine/phosphatidylethanolamine membranes containing various mole fractions of phosphatidylserine was 94–100% at 23°C and 84–88% at 40°C, while the corresponding value for phosphatidylserine/phosphatidylcholine membranes was 74–85% at 23°C and 61–79% at 40°C. Ca2+ concentration required for the phase separation was lower for phosphatidylserine/phosphatidylethanolamine than that for phosphatidylserine/phosphatidylcholine membranes; concentration to cause a half-maximal phase separation was 1.4 · 10?7 M for phosphatidylserine-phosphatidylethanolamine and 1.2 · 10?6 M for phosphatidylserine/phosphatidylcholine membranes. The phase diagram of phosphatidylserine/phosphatidylethanolamine membranes in the presence of Ca2+ was also qualitatively the same as that of phosphatidylserine/phosphatidylcholine except for the different phase transition temperatures of phosphatidylethanolamine (17°C) and phosphatidylcholine (?15°C). These differences were explained in terms of a greater tendency for phosphatidylethanolamine, compared to phosphatidylcholine, to form its own fluid phase separated from the Ca2+-chelated solid-phase phosphatidylserine domain.  相似文献   

16.
Insulin receptor kinase, affinity-purified by adsorption and elution from immobilized insulin, is stimulated 2-3-fold by insulin in detergent solution. Reconstitution of the receptor kinase into leaky vesicles containing phosphatidylcholine and phosphatidylethanolamine (1:1, w/w) by detergent removal on Sephadex G-50 results in the complete loss of receptor kinase sensitivity to activation by insulin. Insulin receptors in these vesicles also exhibit an increase in their apparent affinity for 125I-insulin (Kd = 0.12 nM versus 0.76 nM). Inclusion of 8.3-16.7% phosphatidylserine into the reconstituted vesicles restores 40-50% of the insulin-sensitivity to the receptor kinase. An elevated apparent affinity for 125I-insulin of insulin receptors in vesicles containing phosphatidylcholine and phosphatidylethanolamine is also restored to the value observed in detergent solution by the inclusion of phosphatidylserine in the reconstituted system. The effect of phosphatidylserine on insulin receptor kinase appears specific, because cholesterol, phosphatidylinositol and phosphatidic acid are all unable to restore insulin-sensitivity to the receptor kinase. Autophosphorylation sites on the insulin receptor as analysed by h.p.l.c. of tryptic 32P-labelled receptor phosphopeptides are not different for insulin receptors autophosphorylated in detergent solution or for the reconstituted vesicles in the presence or absence of phosphatidylserine. These data indicate that the phospholipid environment of insulin receptors can modulate its binding and kinase activity, and phosphatidylserine acts to restore insulin-sensitivity to the receptor kinase incorporated into phosphatidylcholine/phosphatidylethanolamine vesicles.  相似文献   

17.
We have investigated the contribution of various phospholipids to membrane fusion induced by divalent cations. Fusion was followed by means of a new fluorescence assay monitoring the mixing of internal aqueous contents of large (0.1 μm diameter) unilamellar liposomes. The rate and extent of fusion induced by Ca2+ in mixed phosphatidylserine/phosphatidylcholine vesicles were lower compared to those in pure phosphatidylserine vesicles. The presence of 50% phosphatidylcholine completely inhibited fusion, although the vesicles aggregated upon Ca2+ addition. When phosphatidylserine was mixed with phosphatidylethanolamine, however, rapid fusion could be induced by Ca2+ even in mixtures that contained only 25% phosphatidylserine. Phosphatidylethanolamine also facilitated fusion by Mg2+ which could not fuse pure phosphatidylserine vesicles. In phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine mixtures, in which the phosphatidylcholine content was kept at 25%, phosphatidylethanolamine could not substitute for phosphatidylserine, and the fusogenic capacity of Mg2+ was abolished by the presence of merely 10% phosphatidylcholine. The initial rate of release of vesicle contents was slower than the rate of fusion in all the mixtures used. The presence of phosphate effected a considerable decrease in the threshold concentration of Ca2+ and also enhanced  相似文献   

18.
The interaction of propranolol with model phospholipid membranes was studied using various experimental techniques. The partition coefficient of propranolol in the negatively charged membranes of vesicles prepared from phosphatidylserine and phosphatidic acid was found to be more than 20-times higher than in neutral phosphatidylcholine membranes. Preferential interaction of propranolol with acidic phospholipid membranes was confirmed using the monolayer compression isotherm technique and the spin-labelling method. Phosphatidylserine monolayers were markedly expanded even at a relatively low drug concentration (5 . 10(-6) M). In contrast, the effect of propranolol on phosphatidylcholine monolayers was much smaller, being detectable only at a higher concentration of the drug (1 . 10(-4) M). Spin-labeling experiments show that propranolol exerts marked ordering effect on bilayers prepared from acidic phospholipids and does not change the order parameter of phosphatidylcholine membranes. The dependence of the propranolol fluorescence spectrum on the polarity of the solvent allowed us to identify the intercalation region of the drug in the membrane. The fluorophore moiety of propranolol was found to be localized in the lipid polar head groups region of the bilayer. The role of electrostatic and hydrophobic effects in propranolol-membrane interaction is discussed and the effect of propranolol on the ordering of phospholipid bilayers is compared with the effects of other anesthetic-like molecules.  相似文献   

19.
We studied the effects of immersion of guinea-pig taenia coli strips in potassium-free media on arachidonate stores and other lipid fractions. Control studies obtained with the strips in Krebs solution showed that greater than 97% of arachidonate was found esterified in phospholipid with the following distribution: phosphatidylethanolamine greater than phosphatidylcholine greater than phosphatidylserine plus phosphatidylinositol. 30 min incubation of the strips with [3H]arachidonate complexed to albumin resulted in incorporation of this isotope into phospholipid and neutral lipid fractions, phosphatidylcholine greater than neutral lipid greater than phosphatidylserine plus phosphatidylinositol greater than phosphatidylethanolamine. 30 min incubations with 32PO4(2-)-resulted in an isotope incorporation into phospholipids, phosphatidylcholine greater than phosphatidylserine plus phosphatidylinositol greater than phosphatidylethanolamine. After 'loading' with [3H]arachidonate and 32P, placing the strips in potassium-free media caused the following: there was an increased release of [3H]arachidonate from the tissue into the bathing solution. [3H]Arachidonate and 32P radioactivity in phosphatidylinositol fell without a change in phosphatidylinositol content. [3H]Arachidonate and 32P radioactivity in other phospholipid fractions was unchanged. Arachidonate specific activity fell and arachidonate content increased in the phosphatidylserine plus phosphatidylinositol fraction. [3]Arachidonate in neutral lipid did not change significantly. We conclude that exposure of taenia coli to potassium-free media activates turnover of phosphatidylinositol, which results in release of arachidonate.  相似文献   

20.
Cytidine, as cytidine 5'-diphosphate choline, is a major precursor in the synthesis of phosphatidylcholine in cell membranes. In the present study, we examined the relationships between extracellular levels of cytidine, the conversion of [14C]choline to [14C]phosphatidylcholine, and the net syntheses of phosphatidylcholine and phosphatidylethanolamine by PC12 cells. The rate at which cytidine (as [3H]cytidine) was incorporated into the PC12 cells followed normal Michaelis-Menten kinetics (Km = 5 microM; Vmax = 12 x 10(-3) mmol/mg of protein/min) when the cytidine concentrations in the medium were below 50 microM; at higher concentrations, intracellular [3H]cytidine nucleotide levels increased linearly. Once inside the cell, cytidine was converted mainly into cytidine triphosphate. In pulse-chase experiments, addition of cytidine to the medium caused a time- and dose-dependent increase (by up to 30%) in the incorporation of [14C]choline into membrane [14C]-phosphatidylcholine. When the PC12 cells were supplemented with both cytidine and choline for 14 h, small but significant elevations (p less than 0.05) were observed in their absolute contents of membrane phosphatidylcholine, phosphatidylethanolamine, and phosphatidylserine, all increasing by 10-15% relative to their levels in cells incubated with choline alone. Exogenous cytidine, acting via cytidine triphosphate, can thus affect the synthesis and levels of cell membrane phospholipids.  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