首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
Liver microsomes were enriched in liposomal acidic lipids by Ca2+-dependent fusion of liposomes at pH 7.0. The extent of fusion was monitored by the transfer of radioactive cholesteryl oleate. The enrichment of membranes in phosphatidylserine inhibited ethanolamine base-exchange, whereas the fusion with phosphatidylinositol inhibited both ethanolamine and serine base-exchange reactions. In contrast, these two phospholipids had scarce effects on choline base-exchange. Phosphatidic acid did not suppress any of the three base-exchange activities. Possible functional implications are discussed.Abbreviations DTT dithiothreitol - HEPES 4-(2-hydroxyethyl)-1-piperazineethansulfonic acid - SHB suerose-HEPES buffer (0.25M sucrose, 3mM HEPES, pH 7.4)  相似文献   

3.
Endogenous substrates (phosphatidylserine, phosphatidylethanolamine and phosphatidylcholine) for the Ca2+-dependent base-exchange reaction were investigated using bovine retinal microsomes. The amounts of the three bases, serine, ethanolamine and choline, released from the membranes and the amount of phosphatidic acid generated in the membranes were measured in the presence of Ca2+ with or without exogenous bases. When the membranes were incubated with Ca2+ alone, the three bases were liberated into the water-soluble fractions accompanied by accumulation of phosphatidic acid, suggesting the presence of Ca2+-dependent phospholipase D-like activity. When an exogenous base was added to the reaction mixture, the liberation of the other two bases increased slightly and the formation of phosphatidic acid decreased markedly. The exogenous base also stimulated the liberation of the same base from prelabeled phospholipids. Accompanying these changes, the exogenous base was incorporated into the membrane phospholipid. With respect to pH profile, time course and metal requirements, both the base incorporation and phospholipase D-like activity were quite similar. The amount of base incorporated generally agreed with both the decreased amount of phosphatidic acid formed and the increased amount of base released. These results suggest that, beside the base-exchange reaction, phospholipase D-like activity plays an important role in Ca2+-dependent base incorporation into bovine retinal membranes.  相似文献   

4.
V79-UF cells were isolated from Chinese hamster V79 cells as a cell line that requires exogenous unsaturated fatty acids for growth. V79-UF cells incorporated arachidonic acid into phospholipids. The molecular species of diacyl phosphatidylcholine and phosphatidylethanolamine containing arachidonic acid comprised 61.4 and 70.5% of the total phospholipid molecular species in total membranes and 58.1 and 64.7% in plasma membrane, respectively. Polyunsaturated molecular species were distributed in a higher amount in the intracellular membranes than in the plasma membrane. No significant difference was seen in the diffusion coefficient between the plasma membranes from cells supplemented with oleic and arachidonic acids in spite of a distinct difference in the degree of unsaturation between the molecular species of these plasma membranes. The amount of cholesterol in the plasma membrane was higher in the cells grown in the presence of arachidonic acid than in those grown in the presence of oleic acid.  相似文献   

5.
The Ca2+ dependent incorporation of [14C]ethanolamine, L-[14C]serine and [14C]choline into phosphatidylethanolamine, phosphatidylserine and phosphatidylcholine, respectively, were investigated in membrane preparations from rat heart. The ethanolamine and serine base-exchange enzyme-catalyzed reactions were associated with the sarcolemma and sarcoplasmic reticulum. There was a 17.2-fold and 6.8-fold enrichment, respectively, of the serine and the ethanolamine base-exchange enzyme activities in the sarcolemma compared to the starting whole homogenate. The sarcoplasmic reticulum was enriched in the ethanolamine and serine base-exchange enzyme activities. The choline base-exchange enzyme activity of all membranes fractions was negligible compared to the ethanolamine or serine base-exchange enzyme activities. The apparent Km for the ethanolamine and serine base-exchange enzyme in sarcolemma was 14 microM and 25 microM, respectively. The pH optimum for these base-exchange activities was 7.5-8.0. There was a dependence upon Ca2+ for these reactions with a 1 or 4 mM concentration required for maximal activity. The properties of the sarcoplasmic reticulum base-exchange enzymes were similar to the sarcolemmal base-exchange enzymes.  相似文献   

6.
In brain, phosphatidylethanolamine can be synthesized from free ethanolamine either by a pathway involving the formation of CDP-ethanolamine and its transfer to diglyceride, or by base-exchange of ethanolamine with existing phospholipids. Although de novo synthesis from serine has also been demonstrated, the metabolic pathway involved is not known. The enzyme phosphatidylserine decarboxylase appears to be involved in the synthesis of much of the phosphatidylethanolamine in liver, but the significance of this route in brain has been challenged. Our in vitro studies demonstrate the existence of phosphatidylserine decarboxylase activity in rat brain and characterize some of its properties. This enzyme is localized in the mitochondrial fraction, whereas the enzymes involved in base-exchange and the cytidine pathway are localized to microsomal membranes. Parallel in vivo studies showed that after the intracranial injection of L-[G-3H]serine, the specific activity of phosphatidylserine was greater in the microsomal fractions than in the mitochondrial fraction, whereas the opposite was true for phosphatidylethanolamine. When L-[U-14C]serine and [1-3H]ethanolamine were simultaneously injected, the 14C/3H ratio in mitochondrial phosphatidylethanolamine was 10 times that in microsomal phosphatidylethanolamine. The results demonstrate that serine is incorporated into the base moiety of phosphatidylethanolamine primarily through the decarboxylation of phosphatidylserine in brain mitochondria. A minimal value of 7% for the contribution of phosphatidylserine decarboxylase to whole-brain phosphatidylethanolamine synthesis can be estimated from the in vivo data.  相似文献   

7.
The increase in light emission of hydroperoxide-supplemented cytochrome c observed on addition of lipid vesicles was related to the degree of unsaturation of the fatty acids of the phospholipids: dipalmitoyl phosphatidylcholine was without effect, whereas dioleoyl phosphatidylcholine and soya-bean phosphatidylcholine enhanced chemiluminescence 2- and 3-fold respectively. Effects on light-emission were similar to those on O2 uptake. The chemiluminescence of the present system was sensitive to cyanide and to the radical trap 2,5-di-t-butylquinol, indicating a catlytic activity of cytochrome c and the presence of free-radical species respectively. Lipid-vesicle enhanced chemiluminescence showed different kinetic behaviours, apparently depending on unsaturation: three phases are described for soya-bean phosphatidylcholine, whereas only one phase was present in mixtures containing dipalmitoyl and dioleoyl phospholipids. Chemiluminescence of lipid vesicles supplemented with cytochrome c and hydroperoxide showed similar kinetic patterns with H2O2 and primary (ethyl) and tertiary (t-butyl and cumene) hydroperoxides. Participation of singlet molecular oxygen, mainly on the phase III of chemiluminescence, is suggested by the increase of light-emission by 1,4-diazabicyclo[2.2.2]-octane as well as by data from spectral analysis.  相似文献   

8.
A study was undertaken on the possible involvement of phospholipids on stereospecific opiate binding to a rat brain membrane fraction comprised mainly of synaptic membranes. The addition of acidic phospholipids such as phosphatidylserine, phosphoinositides, and phosphatidic acid significantly enhanced opiate binding. With the exception of phosphatidylserine, when the acidic phospholipids contained a polyunsaturated acyl group, they were actually inhibitory, along with neutral phospholipids derived from brain. Both the C18:0, C18:1 form (derived from myelin) and the C18:0, C22:6 form of phosphatidylserine (derived from synaptic membranes) produced as much as a 45% enhancement in opiate binding. Unsaturated fatty acids were highly inhibitory, the degree of inhibition being related to the degree of unsaturation. Both phospholipase A and C were inhibitory; and the inhibitory effect of A could not be prevented by albumin or overcome with the addition of phosphatidylserine. With the use of the cross-linking agent, dinitrodifluorobenzene, it could be demonstrated that the phosphatidylserine of synaptic membranes appeared to be preferentially associated with membrane protein. The enhancement of opiate binding by phosphatidylserine diminished with increasing degree of cross-linking.  相似文献   

9.
Mass changes in the various molecular species of phospholipids were determined after stimulation of human platelets with thrombin and collagen. Upon stimulation, every molecular species of phosphatidylinositol and phosphatidylserine was equally hydrolyzed, whereas the molecular species of phosphatidylcholine and diacyl- and alkenylacylphosphatidylethanolamine containing arachidonic acid were selectively hydrolyzed. At low Ca2+ concentrations, which result from mobilization of intracellular Ca2+ stores, phosphatidylinositol, phosphatidylcholine, and diacylphosphatidylethanolamine were hydrolyzed after stimulation with thrombin, whereas only phosphatidylinositol was hydrolyzed with production of thromboxane B2 after stimulation with collagen. At high Ca2+ concentrations, phosphatidylcholine and diacylphosphatidylethanolamine were hydrolyzed after stimulation with collagen, and phosphatidylserine and alkenylacylphosphatidylethanolamine were degraded after stimulation with both thrombin and collagen. [1-14C]Arachidonic acid was heterogeneously incorporated into the individual molecular species of the various phospholipid classes, indicating that the determination of mass is essential for an accurate picture of phospholipid hydrolysis. The data reported here indicate that the Ca2+ concentration affects the differential degradation of phospholipid molecular species in activated human platelets.  相似文献   

10.
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   

11.
The fatty acid composition of phosphatidylcholine (PC) formed by base-exchange was examined in rat brain membranes in vitro. The free choline incorporated into subspecies of PC by this phospholipase-D type activity can be distinguished from that which might enter indirectly via the last enzyme of the de novo pathway for phospholipid biosynthesis, cholinephosphotransferase, by its ionic requirements. Choline base-exchange in lysed synaptosomes is optimal when assayed at extracellular (mM) calcium concentrations and is blocked by magnesium. As much as 40% of the choline incorporated by base-exchange into rat brain membranes was recovered in subspecies of PC, representing no more than 10% of the total PC pool, which contained docosahexaenoic acid (22:6(n-3)). Docosahexaenoic acid is enriched in electrically-excitable membranes and its content in phospholipids of rat and human brain change during early development and increase with age.  相似文献   

12.
The investigation is aimed to study qualitative and quantitative composition of phospholipids, cholesterol content and lipids unsaturation index in plasma membranes of Guerin's carcinoma cells sensitive or resistant to doxorubicin. The comparison of infrared spectra and phospholipids unsaturation index showed that the unsaturation level of fatty acids in plasma membrane from resistant cells was lower than that from sensitive carcinoma cells. 31P-NMR spectroscopy of plasma membranes phospholipids shows the increase of phosphatidylserine and sphingomyeline content in plasma membrane isolated from resistant tumor as compared with sensitive tumor. The levels of phosphatidylcholine and phosphatidylethanolamine were equal in drug-resistant and drug-sensitive carcinoma strains. Changes in plasma membrane from resistant cells result in elevation of plasma membrane microviscosity and phosphatidylserine level increase can suggest the activation of P-glycoprotein-mediated efflux of doxorubicin.  相似文献   

13.
A study was undertaken on the possible involvement of phospholipids on stereospecific opiate binding to a rat brain membrane fraction comprised mainly of synaptic membranes. The addition of acidic phospholipids such as phosphatidylserine, phosphoinositides, and phosphatidic acid significantly enhanced opiate binding. With the exception of phosphatidylserine, when the acidic phospholipids contained a polyunsaturated acyl group, they were actually inhibitory, along with neutral phospholipids derived from brain. Both the C18:0, C18:1 form (derived from myelin) and the C18:0, C22:6 form of phosphatidylserine (derived from synaptic membranes) produced as much as a 45% enhancement in opiate binding. Unsaturated fatty acids were highly inhibitory, the degree of inhibition being related to the degree of unsaturation. Bot phospholipase A and C were inhibitory; and the inhibitory effect of A could not be prevented by albumin or overcome with the addition of phosphatidylserine. With the use of the cross-linking agent, dinitrodifluorobenzene, it could be demonstrated that the phosphatidylserine of synaptic membranes appeared to be preferentially associated with membrane protein. The enhancement of opiate binding by phosphatidylserine diminished with increasing degree of cross-linking.  相似文献   

14.
The effect of synthetic polycations, polyallylamine, and polyethylenimine, on liposomes containing phosphatidylserine was investigated along with that of polylysine and divalent cations. The addition of polycations caused aggregation of sonicated vesicles composed of phosphatidylserine and phosphatidylcholine (molar ratio 1:4) as determined by measuring the turbidity changes. Liposomal turbidity increased 10 times compared with that of control liposomes at charge ratios of polymer/vesicle from 0.23 (polylysine) to 2.5 (linear polyethylenimine), while the turbidity was unchanged by the addition of Ca2+ or Mg2+ at charge ratios up to 500. These polycations also induced intermixing of liposomal membranes as indicated by resonance energy transfer between fluorescent lipids incorporated in lipid bilayers, without inducing drastic permeability changes as determined from the calcein release. Fifty percent intermixing of liposomes (0.05 mM as lipid concentration) was induced by these polycations at charge ratios of around 1.0. However, the highest resonance energy transfer was produced by the addition of polyallylamine, which caused multicycles of membrane intermixing between vesicles. Polycation-induced membrane intermixing and permeability changes of phosphatidylserine liposomes were also investigated. At charge ratios of around 1.0, these polymers caused resonance energy transfer of fluorescent lipids incorporated in separate vesicles; however, polyallylamine and branched polyethylenimine also caused permeability increases of liposomal membranes. Membrane intermixing and permeability changes of phosphatidylserine vesicles induced by polyallylamine were dependent on the polymer/vesicle charge ratio, and were different from those induced by Ca2+ since the latter caused half-maximal membrane intermixing or permeability change of phosphatidylserine vesicles at about 1 mM at the liposomal concentrations investigated.  相似文献   

15.
Incorporation of L-[3-14C]serine into phosphatidylserine (PS) and phosphatidylethanolamine (PE) has been studied in isolated rat hepatocytes. Ethanolamine inhibited the incorporation, indicating competition with serine in the base-exchange reaction. Choline, monomethylethanolamine, dimethylethanolamine and dimethyl-3-aminopropan-1-ol had no such effect. The observed rate of PS biosynthesis corresponded to 7-17 nmol/min per liver at 0.55 mM L-serine. The results indicate that only a small fraction (1/25 to 1/70) of the PS pool equilibrates with the base-exchange enzyme, and that decarboxylation to PE occurs preferentially from this pool. The rate of PS synthesis and decarboxylation can therefore not be calculated by methods which assume random, homogeneous labelling of the total PS pool. The apparent rate of PS decarboxylation increased approx. 4-fold when L-serine increased from 0.5 to 2.25 mM, suggesting that decarboxylation of PS to PE might be regulated by the concentration of L-serine or by the amount of PS present in the hepatocyte cell membranes. Lauric, palmitic, stearic, oleic and linoleic acid decreased the rate of PS synthesis. At 0.5 mM, lauric and palmitic acid were most inhibitory. At 1.0 mM, linoleic acid was the least inhibitory fatty acid. The saturated hexaenoic and saturated tetraenoic species of PS contained 51 and 29%, respectively, of the incorporated L-[3-14C]serine. The combined monoene dienoic/diene dienoic fraction had the highest rate of synthesis judged by its relative specific activity. At 0.9 mM concentration, linoleic acid doubled the relative specific activity of the combined monoene dienoic/diene dienoic fraction of PS. Incorporation of L-[3-14C]serine into molecular species of PE resembled that into PS, both in the absence and presence of linoleic acid, suggesting that the phosphatidylserine decarboxylase (EC 4.1.1.65) has a low specificity towards the fatty acid composition of PS. The results indicate that biosynthesis of PS from L-serine occurs mainly by the base-exchange with only negligible contribution from direct incorporation of phosphatidic acid or diacylglycerol. Furthermore, the deacylation-reacylation pathway seem to contribute only little to the determination of the fatty acid composition of hepatocyte PS. Active PS turnover seems to be confined to a small fraction of the PS pool.  相似文献   

16.
Treatment of bovine thyroid plasma membranes with phospholipase A or C inhibited the stimulation of adenylate cyclase activity by thyroid-stimulating hormone (TSH). In general, basal and NaF-stimulated adenylate cyclase activity was not influenced by such treatment. When plasma membranes were incubated with 1–2 units/ml phospholipase A, subsequent addition of phosphatidylcholine or phosphatidylserine but not phosphatidylethanolamine partially restored TSH stimulation. Phosphatidylcholine was more effective than phosphatidylserine in that it caused greater restoration of the TSH response and smaller amounts of phosphatidylcholine were active. However, when the TSH effect was obliterated by treatment of plasma membranes with 10 units/ml phospholipase A, phospholipids were unable to restore any response to TSH. Lubrol PX, a nonionic detergent, inhibited basal, TSH- and NaF-stimulated adenylate cyclase activities in thyroid plasma membranes. Although phosphatidylcholine partially restored TSH stimulation of adenylate cyclase activity in the presence of Lubrol PX, it did not have a similar effect on the stimulation induced by NaF. These results indicate that phospholipids are probably essential components in the system by which TSH stimulates adenylate cyclase activity in thyroid plasma membranes. The effects do not seem to involve the catalytic activity of adenylate cyclase but the data do not permit a distinction between decreased binding of TSH to its receptor or impairment of the signal from the bound hormone to the enzyme activity.  相似文献   

17.
To identify the specific component(s) in the target membrane involved in fusion of vesicular stomatitis virus (VSV), we examined the interaction of the virus with human erythrocyte membranes with asymmetric and symmetric bilayer distributions of phospholipids. Fusion was monitored spectrofluorometrically by the octadecylrhodamine dequenching assay. Fusion of VSV with lipid-symmetric erythrocyte ghosts was rapid at 37 degrees C and low pH, whereas little or no fusion was observed with lipid-asymmetric ghosts. Conversion of phosphatidylserine in the lipid-symmetric ghost membrane to phosphatidylethanolamine by means of the enzyme phosphatidylserine decarboxylase did not alter the target membrane's susceptibility to VSV fusion. Spin-labeled phospholipid analogues with phosphatidylserine, phosphatidylethanolamine, and phosphatidylcholine headgroups incorporated into the outer leaflet of lipid-asymmetric erythrocytes did not render those membranes fusogenic. Electron spin resonance spectra showed an increased mobility of a phosphatidylcholine spin-label incorporated into the outer leaflet of lipid-symmetric erythrocyte ghosts as compared to that of lipid-asymmetric ghosts. These results indicate that the susceptibility to VSV fusion is not dependent on any particular phospholipid but rather is related to packing characteristics of the target membrane.  相似文献   

18.
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   

19.
Chinese hamster ovary (CHO) cell mutants that required exogenously added phosphatidylserine for cell growth were isolated by using the replica technique with polyester cloth, and three such mutants were characterized. Labeling experiments on intact cells with 32Pi and L-[U-14C]serine revealed that a phosphatidylserine auxotroph, designated as PSA-3, was strikingly defective in phosphatidylserine biosynthesis. When cells were grown for 2 days without phosphatidylserine, the phosphatidylserine content of PSA-3 was about one-third of that of the parent. In extracts of the mutant, the enzymatic activity of the base-exchange reaction of phospholipids with serine producing phosphatidylserine was reduced to 33% of that in the parent; in addition, the activities of base-exchange reactions of phospholipids with choline and ethanolamine in the mutant were also reduced to 1 and 45% of those in the parent, respectively. Furthermore, it was demonstrated that the serine-exchange activity in the parent was inhibited approximately 60% when choline was added to the reaction mixture whereas that in the mutant was not significantly affected. From the results presented here, we conclude the following. There are at least two kinds of serine-exchange enzymes in CHO cells; one (serine-exchange enzyme I) can catalyze the base-exchange reactions of phospholipids with serine, choline, and ethanolamine while the other (serine-exchange enzyme II) does not use the choline as a substrate. Serine-exchange enzyme I, in which mutant PSA-3 is defective, plays a major role in phosphatidylserine biosynthesis in CHO cells. Serine-exchange enzyme I is essential for the growth of CHO cells.  相似文献   

20.
1. The formation of phosphatidylcholine from radioactive precursors was studied in adult rat lung alveolar type II epithelial cells in primary culture. 2. The incorporation of [Me-14C]choline into total lipids and phosphatidylcholine was stimulated by addition of palmitate, whereas the incorporation of [U-14C]glucose into phosphatidylcholine and disaturated phosphatidylcholine was stimulated by addition of choline. Addition of glucose decreased the absolute rate of incorporation of [1(3)-3H]glycerol into total lipids, phosphatidylcholine and disaturated phosphatidylcholine, decreased the percentage [1(3)-3H]glycerol recovered in phosphatidylcholine, but increased the percentage phosphatidylcholine label in the disaturated species. 3. At saturating substrate concentrations, the percentages of phosphatidylcholine radioactivity found in disaturated phosphatidylcholine after incubation with [1-(14)C]acetate (in the presence of glucose) [1-(14)C]palmitate (in the presence of glucose), [Me-14C]choline (in the presence of glucose and palmitate) and [U-14C]glucose (in the presence of choline and palmitate) were 78, 75, 74 and 90%, respectively. 4. Fatty acids stimulated the incorporation of [U-14C]glucose into the glycerol moiety of phosphatidylcholine. The degree of unsaturation of the added fatty acids was reflected in the distribution of [U-14C]glucose label among the different molecular species of phosphatidylcholine. It is suggested that the glucose concentration in the blood as related to the amount of available fatty acids and their degree of unsaturation may be factors governing the synthesis of surfactant lipids.  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