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.
Using large (5-10 microns) vesicles formed in the presence of phospholipids fluorescently labeled on the acyl chain and visualized using a fluorescence microscope, charge-coupled-device camera, and digital image processor, we examined the effects of membrane proteins on phospholipid domain formation. In vesicles composed of phosphatidic acid and phosphatidylcholine, incubation with cytochrome c induced the reorganization of phospholipids into large phosphatidic acid-enriched domains with the exclusion of phosphatidylcholine. Cytochrome c binding was demonstrated to be highest in the phosphatidic acid-enriched domain of the vesicle using the absorbance of the heme moiety for visualization. Both binding of cytochrome c and phospholipid reorganization were blocked by pretreatment of the vesicles with 0.1 M NaCl. The pore forming peptide gramicidin was examined for the effects of an integral protein on domain formation. Initially, gramicidin distributed randomly within the vesicle and showed no phospholipid specificity. Phosphatidic acid domain formation in the presence of 2.0 mM CaCl2 or 100 microM cytochrome c was not affected by the presence of 5 mol % gramicidin within the vesicles. In both cases, gramicidin was preferentially excluded from the phosphatidic acid-enriched domain and became associated with phosphatidylcholine-enriched areas of the vesicle. Thus, cytochrome c caused a major reorganization of both the phospholipids and the proteins in the bilayer.  相似文献   

2.
The purified Ca2+-pumping ATPase of human erythrocyte membranes (Niggli, V., Adunyah, E. S., Penniston, J. T., and Carafoli, E. (1981) J. Biol. Chem. 256, 395-401) can be stimulated, in the absence of calmodulin, by other treatments. 1. A variety of acidic phospholipids (phosphatidylserine, cardiolipin, phosphatidylinositol, and phosphatidic acid) stimulate the Vmax and decrease the Km (Ca2+) of the isolated enzyme to the same extent as calmodulin. Unsaturated fatty acids (oleic and linoleic acid) have the same effect as phospholipids but at lower concentrations. Neutral phospholipids (phosphatidylcholine, sphingomyelin, and phosphatidylethanolamine) have no effect on the enzyme. The minimal proportion of acidic phospholipids in the environment of the enzyme necessary for full stimulation is about 40%. 2. The isolated enzyme, after reconstitution in phosphatidylcholine liposomes in the absence of calmodulin, can be activated by limited proteolysis. The trypsinized enzyme has the same high Vmax and high affinity for Ca2+ of the enzyme in the presence of calmodulin.  相似文献   

3.
A sensitive method which utilizes fluorescence energy transfer to assay Ca2+ -or Mg2+ -mediated fusion of phospholipid vesicles is reported. More than 85% quenching results when phosphatidylserine vesicles labelled with dansyl phosphatidylethanolamine (donor) are fused with vesicles labelled with rhodamine phosphatidylethanolamine (acceptor) in the presence of 5 mM CaCl2 or 10 mM MgCl2. Higher concentrations of divalent cations are required to obtain maximal quenching when phosphatidylserine is partially replaced with phosphatidylethanolamine or phosphatidylcholine. The rate of vesicle fusion is dependent upon the concentrations of both cation and vesicles. Maximum quenching occurs within 5 min using phosphatidylserine vesicles and 5 mM Ca2+, but quenching is incomplete even after 20 h with 0.8--2 mM Ca2+. This probably reflects the heterogeneous size distribution of these vesicles, since the extent of fusion was found to correlated with vesicle size. Binding of antibody to membrane-localized phenobarbital hapten effectively blocks Ca2+ -mediated vesicle fusion. This effect can be inhibited by preincubation of the antibody with phenobarbital. Leakage of tempocholine from intact vesicles induced by 5 mM Ca2+ occurs even when fusion is prevented by bound antibody. This demonstrates that fusion is not a necessary requirement for Ca2+ -induced leakage.  相似文献   

4.
Ionophore A23187, either in the presence or absence of added Ca2+ or Mg2+, caused a marked accumulation of [32P]-phosphatidic acid in pancreatic islets pre-labelled with 32 Pi. A similar effect was observed following the addition of 4 mM Ba2+ ions in the absence of added Ca2+. Neither agent caused a significant modification of labelling in other lipid fractions, although there was a persistent trend towards reduced labelling of phosphatidylcholine and phosphatidylethanolamine. Ionophore A23187 also potentiated the incorporation of 3H-glycerol into phosphatidic acid and reduced the incorporation of this precursor into phosphatidylcholine. In islets pre-labelled with 3H-glycerol and subsequently exposed to A23187 or Ba2+, no significant changes were observed in label associated with either phospholipids or neutral glycerolipids. These results suggest that ionophore A23187 and Ba2+ ions can divert the synthesis of phospholipids resulting in increased formation of phosphatidic acid at the expense of non-acidic phospholipids, principally phosphatidylcholine. We tentatively suggest that this effect may be the result of inhibition by Ca2+ of the breakdown of phosphatidic acid to diglyceride, an enzymic step which may regulate the relative amounts of acidic and neutral phospholipids.  相似文献   

5.
A phospholipid transfer protein from yeast (Daum, G. and Paltauf, F. (1984) Biochim. Biophys. Acta 794, 385-391) was 2800-fold enriched by an improved procedure. The specificity of this transfer protein and the influence of membrane properties of acceptor vesicles (lipid composition, charge, fluidity) on the transfer activity were determined in vitro using pyrene-labeled phospholipids. The yeast transfer protein forms a complex with phosphatidylinositol or phosphatidylcholine, respectively, and transfers these two phospholipids between biological and/or artificial membranes. The transfer rate for phosphatidylinositol is 19-fold higher than for phosphatidylcholine as determined with 1:8 mixtures of phosphatidylinositol and phosphatidylcholine in donor and acceptor membrane vesicles. If acceptor membranes consist only of non-transferable phospholipids, e.g., phosphatidylethanolamine, a moderate but significant net transfer of phosphatidylcholine occurs. Phosphatidylcholine transfer is inhibited to a variable extent by negatively charged phospholipids and by fatty acids. Differences in the accessibility of the charged groups of lipids to the transfer protein might account for the different inhibitory effects, which occur in the order phosphatidylserine which is greater than phosphatidylglycerol which is greater than phosphatidylinositol which is greater than cardiolipin which is greater than phosphatidic acid which is greater than fatty acids. Although mitochondrial membranes contain high amounts of negatively charged phospholipids, they serve effectively as acceptor membranes, whereas transfer to vesicles prepared from total mitochondrial lipids is essentially zero. Ergosterol reduces the transfer rate, probably by decreasing membrane fluidity. This notion is supported by data obtained with dipalmitoyl phosphatidylcholine as acceptor vesicle component; in this case the transfer rate is significantly reduced below the phase transition temperature of the phospholipid.  相似文献   

6.
Fluorescent derivatives of a phosphatidylglycerol, phosphatidylserine, phosphatidic acid, phosphatidylcholine, phosphatidylethanolamine, and diacylglycerol have been studied to establish the effect of different polar headgroups on the mechanism and kinetics of spontaneous phospholipid transfer between recombinants of human plasma apolipoprotein A-II and dimyristoylphosphatidylcholine. The fluorescent lipids are all 1-myristoyl-2-[9-(1-pyrenyl)nonanoyl] glycerides. The transfer of the lipids is a first order process where the rate is independent of the concentration over a 50 fold range of the acceptor recombinants. These results are consistent with the lipids transferring as monomers being a water-soluble intermediate. The rate of transfer of the different phospholipids are slightly slower than phosphatidylcholine, with that of phosphatidylethanolamine being about 4 times slower. The transfer of phospholipids with a titratable headgroup is pH-dependent. The difference in the rates and pH dependence may be a function of the interactions (hydrogen bonding) between polar headgroups. The rate of transfer of the diacylglycerol is 20 times slower than phosphatidylcholine, but its activation energy (21 kcal/mol) is only 2 to 3 kcal less than most of the phospholipids (23 kcal/mol). These results suggest that the rate and activation energy for the spontaneous transfer of phospholipids can be predicted to a first approximation on the basis of its hydrophobic content, irrespective of the pH or identity of the polar headgroup.  相似文献   

7.
Purified Acetylcholine Receptor (AcChR) from Torpedo has been reconstituted at low (approximately 1:3500) and high (approximately 1:560) protein to phospholipid molar ratios into vesicles containing egg phosphatidylcholine, cholesterol, and different dimyristoyl phospholipids (dimyristoyl phosphatidylcholine, phosphatidylserine, phosphatidylglycerol and phosphatidic acid) as probes to explore the effects of the protein on phospholipid organization by differential scanning calorimetry, infrared, and fluorescence spectroscopy. All the experimental results indicate that the presence of the AcChR protein, even at the lower protein to phospholipid molar ratio, directs lateral phase separation of the monoanionic phosphoryl form of the phosphatidic acid probe, causing the formation of specific phosphatidic acid-rich lipid domains that become segregated from the bulk lipids and whose extent (phosphatidic acid sequestered into the domain, out of the total population in the vesicle) is protein-dependent. Furthermore, fluorescence energy transfer using the protein tryptophan residues as energy donors and the fluorescence probes trans-parinaric acid or diphenylhexatriene as acceptors, establishes that the AcChR is included in the domain. Other dimyristoyl phospholipid probes (phosphatidylcholine, phosphatidylserine, phosphatidylglycerol) under identical conditions could not mimic the protein-induced domain formation observed with the phosphatidic acid probe and result in ideal mixing of all lipid components in the reconstituted vesicles. Likewise, in the absence of protein, all the phospholipid probes, including phosphatidic acid, exhibit ideal mixing behavior. Since phosphatidic acid and cholesterol have been implicated in functional modulation of the reconstituted AcChR, it is suggested that such a specific modulatory role could be mediated by domain segregation of the relevant lipid classes.  相似文献   

8.
B Mitra  G G Hammes 《Biochemistry》1989,28(7):3063-3069
Fluorescence resonance energy-transfer measurements were made on the membrane-bound chloroplast coupling factor. The distances from the N,N'-dicyclohexylcarbodiimide-binding site on the membrane-bound portion of the enzyme (CF0) to the vesicle surface and to two sulfhydryl sites on the gamma-polypeptide were determined. The dicyclohexylcarbodiimide-binding site was labeled with the fluorescent species N-cyclohexyl-N'-pyrenylcarbodiimide. The vesicle surface was labeled with N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine. Steady-state energy transfer between the fluorescent-labeled enzyme (energy donor) and varying concentrations of the ethanolamine derivative (energy acceptor) indicated that the distance of closest approach between the energy donor and the outer vesicle surface is 16-24 A. Two specific sites on the gamma-polypeptide were reacted with a coumarinylmaleimide derivative; one is a sulfhydryl that can be labeled only on the thylakoids under energized conditions (the "light" site), while the other is the disulfide site that regulates enzymatic activity. Energy-transfer measurements utilizing steady-state fluorescence and fluorescence lifetime methods indicated that the dicyclohexylcarbodiimide site is approximately 41 A from the light site and approximately 50 A from the gamma-disulfide site. These distances are used to extend the current structural model of the chloroplast coupling factor.  相似文献   

9.
The effects of Na+, other cations and the neurotransmitters, acetylcholine and norepinephrine on 32Pi incorporation into phospholipids of the rabbit iris smooth muscle were investigated [1]. The basal 32P-labelling of phospholipids including phosphatidic acid, phosphatidylinositol, phosphatidylcholine, phosphatidylethanolamine and the polyphosphoinositides increased with Na+ concentration [2]. The neurotransmitter-stimulated 32P labelling of phosphatidic acid, phosphatidylinositol and phosphatidylcholine is dependent on the presence of extracellular Na+ [3]. The monovalent cation requirement for Na+ specific. Of the monovalent cations Li+, NH+4, K+, Choline+ and Tris, only Li+ partially substituted for Na+ [4]. A significant decrease in 32P labelling of phospholipids in response to acetylcholine was observed when Ca2+ and/or K+ were added to an isoosmotic medium deficient of Na+ [5]. Ouabain, which blocks the Na+-pump, inhibited the basal 32Pi incorporation into phosphatidylcholine and the acetylcholine-stimulated 32P labelling of phosphatidic acid, phosphatidylinositol and phosphatidylcholine [6]. It was suggested that phosphoinositide breakdown is associated with Ca2+ influx as we have previously reported (Akhtar, R.A. and Abdel-Latif, A.A. (1978) J. Pharmacol. Exp. Ther. 204, 655-668) and that the enhanced 32P-labelling of phosphoinositides could be associated with Na+ outflux, via the Na+-pump mechanism.  相似文献   

10.
Human platelets are routinely stored for 5 days prior to transfusion, but they deteriorate during storage. Since very little information is available concerning the effect of storage on platelet phospholipid metabolism, the biosynthesis and remodelling of platelet phospholipids were studied. Platelets were incubated separately with [14C]glycerol, [14C]arachidonic acid, or a mixture of [14C]glycerol and [3H]arachidonic acid, and stored in a platelet storage medium at 22 degrees C. Maximum glycerol uptake (20%) was attained after 6 h. [14C]Glycerol was incorporated into phosphatidylcholine, phosphatidylethanolamine, and phosphatidylinositol, and to a much lesser extent phosphatidylserine, under storage conditions for 5 days. The distribution of the initial arachidonic acid uptake was not as would be expected based on the molar composition of endogenous phospholipids. The arachidonic acid (75%) which was taken up within 10 min of incubation distributed 55% into the phosphatidylcholine and only 14% into the phosphatidylethanolamine; the molar composition is actually 18% phosphatidylcholine and 47% phosphatidylethanolamine. During storage, there was a continuous transfer of the radiolabelled arachidonic from phosphatidylcholine to phosphatidylethanolamine until, after 5 days, the distribution of arachidonic acid was identical to the endogenous distribution. In contrast, no change in the glycerol incorporation pattern was detected during storage. This suggested that the mechanism for arachidonic acid redistribution was not through exchange of polar head groups, but through acyl transfer of arachidonic acid from phosphatidylcholine to phosphatidylethanolamine.  相似文献   

11.
The ability of oligo- and polymers of the basic amino acids L-lysine, L-arginine, L-histidine and L-ornithine to induce lipid intermixing and membrane fusion among vesicles containing various anionic phospholipids has been investigated. Among vesicle consisting of either phosphatidylinositol or mixtures of phosphatidic acid and phosphatidylethanolamine rapid and extensive lipid intermixing, but not complete fusion, was induced at neutral pH by poly-L-ornithine or L-lysine peptides of five or more residues. When phosphatidylcholine was included in the vesicles, the lipid intermixing was severely inhibited. Such lipid intermixing was also much less pronounced among phosphatidylserine vesicles. Poly-L-arginine provoked considerable leakage from the various anionic vesicles and caused significantly less lipid intermixing than L-lysine peptides at neutral pH. When the addition of basic amino acid polymer was followed by acidification to pH 5-6, vesicle fusion was induced. Fusion was more pronounced among vesicles containing phosphatidylserine or phosphatidic acid than among those containing phosphatidylinositol, and occurred also with vesicles whose composition resembles that of cellular membranes (i.e., phosphatidylcholine/phosphatidylethanolamine/phosphatidylserine, 50:30:20, by mol). Liposomes with this composition are resistant to fusion by Ca2+ or by acidification after lectin-mediated contact. The tight interaction among vesicles at neutral pH, resulting in lipid intermixing, does not seem to be necessary for the fusion occurring after acidification, but the basic peptides nevertheless appear to play a more active role in the fusion process than simply bringing the vesicles in contact. However, protonation of the polymer side chains and transformation of the polymer into a polycation does not explain the need for acidification, since the pH-dependence was quite similar for poly(L-histidine)- and poly(L-lysine)-mediated fusion.  相似文献   

12.
A membrane preparation from porcine platelets catalyzed the hydrolysis of [2-3H]glycerol-labeled lysophosphatidylinositol to form monoacylglycerol and inositol phosphates. The hydrolysis was optimal at pH 9. The addition of Ca2+ did not enhance the hydrolysis, but the enzyme was inhibited completely by EGTA. The EGTA-inactivated enzyme was partially reactivated by Ca2+; Mn2+, Mg2+, and Zn2+ were much less effective or ineffective for the reactivation. The phospholipase C was apparently specific for lysophosphatidylinositol; phosphatidylinositol, phosphatidylcholine, phosphatidylethanolamine, lysophosphatidylcholine, lysophosphatidylethanolamine, phosphatidic acid, and lysophosphatidic acid were not hydrolyzed at significant rates under the conditions used. Phospholipase C with these properties has not been reported previously.  相似文献   

13.
Hydrolysis of lipid mixtures by rat hepatic lipase   总被引:1,自引:0,他引:1  
The hydrolysis of phospholipid mixtures by purified rat hepatic lipase, also known as hepatic triglyceride lipase, was studied in a Triton X-100/lipid mixed micellar system. Column chromatography of the mixed micelles showed elution of Triton X-100 and binary lipid mixtures of phosphatidic acid, phosphatidylcholine, phosphatidylethanolamine and phosphatidylserine as a single peak. This indicated that the mixed micelles were homogenous and contained all components in the designated molar ratios. The molar ratio of Triton X-100 to lipid was kept constant at 4 to 1. Labeling one lipid with 3H and the other lipid with 14C enabled us to determine the hydrolysis of both components of these binary lipid mixed micelles. We found that the hydrolysis of phosphatidylcholine was activated by the inclusion of small amounts of phosphatidic acid (2.5-fold), phosphatidylethanolamine (1.5-fold) or phosphatidylserine (1.4-fold). The maximal activation of phosphatidylcholine hydrolysis was observed when 5 mol% of phosphatidylethanolamine, 7.5 mol% phosphatidic acid or 5 mol% phosphatidylserine was added to Triton X-100 mixed micelles. The hydrolysis of phosphatidic acid was activated 30%, and that of phosphatidylserine was inhibited 30% when the molar proportion of phosphatidylcholine was less than 50 mol%. The hydrolysis of phosphatidylethanolamine was slightly activated when the mol% of phosphatidylcholine was below 5. The hydrolysis of phosphatidylserine was inhibited by phosphatidylethanolamine when the mol% of the latter was 50 or less whereas phosphatidylethanolamine hydrolysis was not affected by phosphatidylserine. Under the conditions used sphingomyelin and cholesterol did not have a significant effect on the hydrolysis of the phospholipids studied. In agreement with our previous study (Kucera et al. (1988) J. Biol. Chem. 263, 1920-1928) these studies show that the phospholipid polar head group is an important factor which influences the action of hepatic lipase and that the interfacial properties of the substrate play a role in the expression of the activity of this enzyme. The molar ratios of phosphatidic acid, phosphatidylethanolamine and phosphatidylserine which activated phosphatidylcholine hydrolysis correspond closely to the molar ratios of these lipids found in the surface lipid film of lipoproteins e.g., high density lipoproteins.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
Human erythrocytes were cholesterol-depleted (5-25%) by incubation with phosphatidylcholine vesicles in media containing Ca2+ at different concentrations (0, 28 nM, 5 microM or 1 mM). After removal of the vesicles, the cells were reincubated with [32P]phosphate in the same media. Control (incubated in buffer alone) and cholesterol-maintained erythrocytes (incubated with cholesterol/phosphatidylcholine vesicles) were treated similarly. Cholesterol depletion induced the conversion of the cells into stomatocytes III and spherostomatocytes and decreased the turnover rate of phosphatidylinositol phosphate and of phosphatidylinositol bisphosphate. None of these effects were observed in cholesterol-maintained cells. In cholesterol-depleted cells, they occurred without changes in the ATP specific activity or in the polyphosphoinositide concentrations. Moreover, these modifications of shape and of lipid metabolism were proportional to the extent of the cholesterol depletion and were independent of the external Ca2+ concentration. In contrast, other effects of cholesterol depletion, a decrease in the turnover rate of phosphatidic acid, a decrease in diacylglycerol and in phosphatidic acid concentrations were dependent on the external Ca2+ concentration. Thus it appears that the shape change was not correlated with a change in the concentrations of these phospholipids or of diacylglycerol and therefore cannot be explained by a bilayer couple mechanism involving these phospholipids. However, the spherostomatocytic transformation was correlated with the decrease in the turnover rate of the polyphosphoinositides, but not with the turnover rate of phosphatidic acid, suggesting a role for the turnover of the polyphosphoinositides in the maintenance of the erythrocyte shape.  相似文献   

15.
Matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry is an established tool for the analysis of proteins, whereas it gained by far less interest in the field of lipid analysis. This method works well with phospholipids as well as organic cell extracts and provides high sensitivity and reproducibility. The aim of the present paper is to extend our previous studies to the analysis of lysophospholipids and phospholipid mixtures. To study the suitability of MALDI-TOF mass spectrometry for the analysis of lysophospholipids, different phospholipids like phosphatidylcholine, phosphatidylethanolamine, phosphatidylserine, phosphatidic acid, and phosphatidylinositol as well as their mixtures were digested with phospholipase A(2). Positive and negative ion mass spectra of all phospholipids before and after digestion were recorded. In all these cases, the molecular ions of the expected digestion products could be detected and only a very small extent of further fragmentation was observed. On the other hand, spectra of phospholipid mixtures containing phosphatidylcholine were strongly dominated by phosphatidylcholine and lysophosphatidylcholine signals, which prevented the detection of further phospholipids even if those lipids were present in comparable amounts. This is of paramount interest for the analysis of tissue and cell extracts.  相似文献   

16.
Once brain ischemia was induced in the gerbil cerebral fronto-parietal cortex, serial changes occurred in energy metabolites and various lipids. The amounts of inositol-containing phospholipids began to decrease immediately after energy failure, followed by an increase in the amount of 1,2-diacylglycerol with a subsequent liberation of arachidonic acid and other free fatty acids. The fatty acid compositions of inositol-containing phospholipids, of 1,2-diacylglycerols produced by ischemia, and of free fatty acids liberated during ischemia were quite similar. The amount of stearic acid liberated was much larger than that of arachidonic acid between 30 s and 1 min of ischemia. On the other hand, there was no significant decrease in the amount of the other phospholipids except for phosphatidic acid. Furthermore, there was also no change in the fatty acid composition of phosphatidylcholine or phosphatidylethanolamine throughout 15 min of ischemia. The amount of cytidine-monophosphate reached a peak (36.7 nmol/g wet wt) at 2 min of ischemia. These results indicated that arachidonic acid was predominantly liberated from inositol-containing phospholipids by phospholipase C, and by the diglyceride lipase and monoglyceride lipase system rather than from phosphatidylcholine or phosphatidylethanolamine by phospholipase A2 or plasmalogenase or choline phosphotransferase during the early period of ischemia.  相似文献   

17.
The metabolism of phospholipids in the brown alga, Fucus serratus was studied. The major phospholipids of this alga are phosphatidylethanolamine, phosphatidylglycerol, phosphatidylinositol, cardiolipin and phosphatidylcholine. When the time-course of labelling of the lipids from [32P] orthophosphate was studied, total labelling was approximately linear for 8 hr. All the major classes of phospholipid were labelled. The extent and pattern of labelling were not affected by the presence of proteins synthesis inhibitors phosphatidic acid was highly labelled at short time intervals. Phosphatidylcholine was relatively poorly labelled. The extent and pattern of labelling were not affected by the presence of protein synthesis inhibitors indicating that the enzymes involved in phospholipid synthesis have a rather slow turnover. Incorporation of radioactivity into phosphatidylglycerol was stimulated significantly by light.  相似文献   

18.
Rat platelet homogenates are able to catalyze CoA-mediated, ATP-independent transfer of arachidonic acid from platelet phospholipids to added lysophospholipids. Homogenates of platelets prelabelled with radioactive arachidonic or oleic acid were incubated in the presence of CoA and various lysophospholipids. Transfer observed with arachidonic acid-labelled platelets was dependent on the lysophospholipid added. When 1-alkenyl- or 1-acyllysophosphatidylethanolamine was used, there was a more efficient arachidonyl transfer from phosphatidylcholine than from phosphatidylinositol to the phosphatidylethanolamine fraction. Lysophosphatidylserine also accepted arachidonyl from phosphatidylcholine. Addition of lysophosphatidylcholine resulted in a decrease in the labelling of phosphatidylinositol and to a lesser extent of phosphatidylethanolamine with concomitant transfer to phosphatidylcholine. Lysophosphatidylinositol and lysophosphatic acid did not act as substrate for this transfer reaction. Free, non-radioactive arachidonic acid did not compete for the labelled arachidonic acid transfer. This pathway may play a major role in the synthesis of arachidonyl species of phosphatidylethanolamine and phosphatidylserine and for the arachidonyl transfer to the phosphatidylethanolamine plasmologen in stimulated platelets.  相似文献   

19.
Ca2+-translocating activities of phosphatidylinositol, diacylglycerol and phosphatidic acid were investigated in phosphatidylcholine liposomes. Using a fluorescent indicator of Ca2+ concentration, quin-2, release of encapsulated Ca2+ from egg yolk phosphatidylcholine liposomes containing 2 mol% of one of these lipids was measured at 37 degrees C. The rate of Ca2+ translocation across the liposomal membrane mediated by phosphatidic acid was about 3-fold larger than those mediated by phosphatidylinositol and diacylglycerol. The result implies that phosphatidic acid has Ca2+-ionophore activity in the agonist dependent metabolism of inositol phospholipids. The ionophoretic activity depended on the degree of unsaturation of the fatty acyl chains. The Ca2+ translocation rate was smallest in dipalmitoylphosphatidic acid, and it increased in the order of dioleoyl-, dilinoleoyl- and dilinolenoyl-phosphatidic acid. Ca2+ mobilization of a stimulated cell is discussed in the light of Ca2+-ionophore activity of phosphatidic acid converted from inositol phospholipids.  相似文献   

20.
The effect of phosphorylation of pre interleukin 1 alpha (IL 1 alpha) on its association with various phospholipids was investigated. We prepared genetically engineered truncated human pre IL 1 alpha (residues 64 to 271) and phosphorylated this pre IL 1 alpha in vitro by using the catalytic subunit of cAMP-dependent protein kinase. Phosphorylated truncated pre IL 1 alpha selectively binds to acidic phospholipids including phosphatidic acid, phosphatidylserine, and phosphatidylinositol, but not to other phospholipids (phosphatidylcholine and phosphatidylethanolamine). This binding required divalent cations: Ca2+ or Mn2+, but not Mg2+. In order to obtain half-maximal binding of pre IL 1 alpha to phosphatidic acid or phosphatidylserine, Ca2+ between 5 and 100 microM was required. Unphosphorylated pre IL 1 alpha did not bind to phosphatidylserine, indicating that phosphorylation is required for this binding. Phosphorylated pre IL 1 alpha did not bind to intact peripheral blood mononuclear cells irrespective of lipopolysaccharide stimulation, but did bind to membrane vesicles prepared from these cells in the presence of calcium. Furthermore, phosphorylated pre IL 1 alpha bound only to inside-out ghosts, but not right-side-out ghosts, prepared from human red blood cells. Taken together, these data suggest that phosphorylated pre IL 1 alpha binds to the inner surface of plasma membrane in a Ca2(+)- and phospholipid-dependent manner.  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