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.
1. In order to better understand ligand-induced redistribution of membrane receptors and lymphocyte activation in ectothermic vertebrates, flow cytometry was used to monitor the effects of both in vivo acclimation temperature and in vitro assay temperatures on the kinetics of monoclonal antibody-induced membrane immunoglobulin (mIg) capping on channel catfish lymphocytes. 2. It was observed that the kinetics of mIg capping were dependent on in vitro assay temperatures, in vivo acclimation temperatures, and the length of time of in vivo acclimation. In the latter situation in vivo acclimation of fish to 27, 22 and 17 degrees C was considered complete after 3 weeks, while acclimation to 12 degrees C required a minimum of 5 weeks. 3. The energies of activation required for mIg capping ranged from 33 to 24 kcal/mol; lower energies of activation were observed with lower temperature acclimation. 4. It was also noted that the lower energies of activation were associated with concomitant decreases in cellular phospholipid saturated/unsaturated fatty acid ratios. 5. It appears that channel catfish B cell mIg capping, presumably a requisite for immune function, can be significantly affected by environmental temperatures; most likely such effects are attributable to changes in plasma membrane viscosities.  相似文献   

2.
M A Cuchens  T M Buttke 《Cytometry》1984,5(6):601-609
A flow cytometric procedure has been developed for performing kinetic studies on the capping of membrane immunoglobulin (mIg) on B lymphocytes. Mouse B cells were stained with fluorescein-conjugated antimouse-Ig antisera and subjected to pulse-shape (width, peak, and area) analyses prior to, during, and after ligand-induced redistribution of mIg. It was found that ring-stained, patched, and capped cells could be discriminated based on the width of the electronic signal curve generated as the cells passed through the laser beam. Additionally, endocytosis and or shedding of the cap could be correlated with a change in the area under the curve. Using these two parameters (width and area), the effects of temperature, cross-linking, and several pharmacological agents on the capping process were examined. Through the use of flow cytometry, the inhibitory effects of various perturbants could be localized to discrete stages of the capping process.  相似文献   

3.
Stabilization of red cells against hypotonic haemolysis by cis-trans isomeric free C18 fatty acids occurs with pronounced specificity which is strongly temperature-dependent, but in a distinctly different manner for the two configurational isomers. Oleic acid (cis-18:1) stabilizes very efficiently at 0 degrees C, even at the highest concentrations. Elaidic acid (trans-18:1) causes neither stabilization nor haemolysis at this temperature. At room temperature (23 degrees C), elaidic acid acquires the ability to protect, without turning haemolytic at high concentrations. At 37 degrees C elaidic acid also becomes haemolytic. The protecting effect of oleic acid at 0 degrees C is the result of a rapid reaction. The characteristic, temperature-dependent specificity of cis-trans isomeric C18 fatty acid interaction with the red cell membrane appears to be a general phenomenon, since it was observed alike with erythrocytes of different species.  相似文献   

4.
Human red cell membrane Ca2+-stimulatable, Mg2+-dependent adenosine triphosphatase (Ca2+-ATPase) activity and its response to thyroid hormone have been studied following exposure of membranes in vitro to specific long-chain fatty acids. Basal enzyme activity (no added thyroid hormone) was significantly decreased by additions of 10(-9)-10(-4) M-stearic (18:0) and oleic (18:1 cis-9) acids. Methyl oleate and elaidic (18:1 trans-9), palmitic (16:0) and lauric (12:0) acids at 10(-6) and 10(-4) M were not inhibitory, nor were arachidonic (20:4) and linolenic (18:3) acids. Myristic acid (14:0) was inhibitory only at 10(-4) M. Thus, chain length of 18 carbon atoms and anionic charge were the principal determinants of inhibitory activity. Introduction of a cis-9 double bond (oleic acid) did not alter the inhibitory activity of the 18-carbon moiety (stearic acid), but the trans-9 elaidic acid did not cause enzyme inhibition. While the predominant effect of fatty acids on erythrocyte Ca2+-ATPase in situ is inhibition of basal activity, elaidic, linoleic (18:2) and palmitoleic (16:1) acids at 10(-6) and 10(-4) M stimulated the enzyme. Methyl elaidate was not stimulatory. These structure-activity relationships differ from those described for fatty acids and purified red cell Ca2+-ATPase reconstituted in liposomes. Thyroid hormone stimulation of Ca2+-ATPase was significantly decreased by stearic and oleic acids (10(-9)-10(-4) M), but also by elaidic, linoleic, palmitoleic and myristic acids. Arachidonic, palmitic and lauric acids were ineffective, as were the methyl esters of oleic and elaidic acids. Thus, inhibition of the iodothyronine effect on Ca2+-ATPase by fatty acids has similar, but not identical, structure-activity relationships to those for basal enzyme activity. To examine mechanisms for these fatty acid effects, we studied the action of oleic and stearic acids on responsiveness of the enzyme to purified calmodulin, the Ca2+-binding activator protein for Ca2+-ATPase. Oleic and stearic acids (10(-9)-10(-4) M) progressively inhibited, but did not abolish, enzyme stimulation by calmodulin (10(-9) M). Double-reciprocal analysis of the effect of oleic acid on calmodulin stimulation indicated noncompetitive inhibition. Addition of calmodulin to membranes in the presence of equimolar oleic acid restored basal enzyme activity. Oleic acid also reduced 125I-calmodulin binding to membranes, but had no effect on the binding of [125I]T4 by ghosts. The mechanism of the decrease by long chain fatty acids of Ca2+-ATPase activity in situ in human red cell ghosts thus is calmodulin-dependent and involves reduction in membrane binding of calmodulin.  相似文献   

5.
Growth of Japanese encephalitis virus (JEV) in BHK-21 cells was stimulated in the presence of 20 to 40 mug of the sodium salt of oleic acid (cis-9-octadecenoic acid, 9-18:1) per ml supplemented in Waymouth medium. The stimulatory effect of the salt was highest when 9-18:1 was added after adsorption of the virus. Study of the effect of other fatty acids on growth of JEV showed the following results: the longer the chain length of the saturated fatty acid salt, the higher the stimulatory effect on viral growth. In contrast, polyunsaturated fatty acids had an inhibitory effect on viral growth. The effect of isomeric cis-octadecenoic acids on viral growth was variable, depending upon the position of the double bond. The cis-6-octadecenoic acid had the highest inhibitory effect on growth of JEV compared to other isomeric octadecenoic acids. The sodium salt of (1-14C) cis-9-octadecenoic acid (9-18:1, 20 mug/ml) was rapidly incorporated into control and JEV-infected cells. Specific radioactivity in phosphatidylcholine dropped 12 to 24 h after virus inoculation, whereas synthesis of phosphatidylethanolamine increased 12 to 24 h after virus inoculation in infected cells compared to uninfected cells. Results from these studies suggest that phospholipid metabolism of infected cells is markedly changed, which can be associated with altered fatty acid metabolism when using labeled 9-18:1 fatty acid as a marker.  相似文献   

6.
The origin and metabolism of octadecenoic acid (18 : 1) was examined in intact Novikoff rat hepatoma cells by using labeled precursors and two isomeric octadecenoic acids which differed in their abilities to stimulate cell growth in a serum-free medium. The isomers (ci-6-18 : 1 and cis-9-18 : 1) were measured in the cellular lipid by ozonolysis and reduction of the ozonides. The results indicate that the 18 : 1 fatty acid accumulated in the cell lipid by uptake of the preformed acid from the medium. The cis-9-18 : 1 to 16 : 1 and 20 : 1 fatty acids by chain shortening and chain elongation. Both isomers inhibited de novo fatty acid synthesis from acetate by cells suspended in a serum-free medium. The isomers did not exert coordinate control of both fatty acid and cholesterol biosynthesis in the Novikoff cells.  相似文献   

7.
Structural determination of polyunsaturated fatty acids by gas chromatography-mass spectrometry (GC-MS) requires currently the use of nitrogen containing derivatives such as picolinyl esters, 4,4-dimethyloxazoline or pyrrolidides derivatives. The derivatization is required in most cases to obtain low energy fragmentation that allows accurate location of the double bonds. In the present work, the following metabolites of rumelenic (cis-9,trans-11,cis-15 18:3) acid, from rat livers, were identified: cis-8,cis-11,trans-13,cis-17 20:4, cis-5,cis-8,cis-11,trans-13,cis-17 20:5, cis-7,cis-10,cis-13,trans-15,cis-19 22:5, and cis-4,cis-7,cis-10,cis-13,trans-15,cis-19 22:6 acids by GC-MS as their 4,4-dimethyloxazoline and methyl esters derivatives. Specific fragmentation of the methyl ester derivatives revealed some similarity with their corresponding DMOX derivatives. Indeed, intense ion fragments at m/z=M+-69, corresponding to a cleavage at the center of a bis-methylene interrupted double bond system were observed for all identified metabolites. Moreover, intense ion fragments at m/z=M+-136, corresponding to allylic cleavage of the n-12 double bonds were observed for the C20:5, C22:5, C22:6 acid metabolites. For the long chain polyunsaturated fatty acids from the rumelenic metabolism, we showed that single methyl esters derivatives might be used for both usual quantification by GC-FID and identification by GC-MS.  相似文献   

8.
We have previously reported that fetal rabbit lung tissue in organ culture produces a lamellar body material (pulmonary surfactant) with a lower percentage of disaturated phosphatidylcholine than is typically found in rabbit lung in vivo (Longmuir, K.J., C. Resele-Tiden, and L. Sykes. 1985. Biochim. Biophys. Acta. 833: 135-143). This investigation was conducted to identify all fatty acids present in the lamellar body phosphatidylcholine, and to determine whether the low level of disaturated phosphatidylcholine is due to excessive unsaturated fatty acid at position sn-1, sn-2, or both. Fetal rabbit lung tissue, 23 days gestation, was maintained in culture for 7 days in defined (serum-free) medium. Phospholipids were labeled in culture with [1-14C]acetate or [U-14C]glycerol (to follow de novo fatty acid biosynthesis), or with [1-14C]palmitic acid (to follow incorporation of exogenously supplied fatty acid). Radiolabeled fatty acid methyl esters obtained from lamellar body phosphatidylcholine were first separated by reverse-phase thin-layer chromatography (TLC) into two fractions of 1) 14:0 + 16:1 and 2) 16:0 + 18:1. Complete separation of the individual saturated and monoenoic fatty acids was achieved by silver nitrate TLC of the two fractions. Monoenoic fatty acid double bond position was determined by permanganate-periodate oxidation followed by HPLC of the carboxylic acid phenacyl esters. Lamellar body phosphatidylcholine contained four monoenoic fatty acids: 1) palmitoleic acid, 16:1 cis-9; 2) oleic acid, 18:1 cis-9; 3) cis-vaccenic acid, 18:1 cis-11; and 4) 6-hexadecenoic acid, 16:1 cis-6. In addition, 8-octadecenoic acid, 18:1 cis-8, was found in the fatty acids of the tissue homogenate. The abnormally low disaturated phosphatidylcholine content in lamellar body material was the result of abnormally high levels of monoenoic fatty acid (principally 16:1 cis-9) found at position sn-2. Position sn-1 contained normal levels of saturated fatty acid. The biosynthesis of the unusual n-10 fatty acids was observed from the start of culture throughout the entire 7-day culture period, and was observed in incubations of tissue slices of day 23 fetal rabbit lung. This is the first report of the biosynthesis of n-10 fatty acids (16:1 cis-6 and 18:1 cis-8) in a mammalian tissue other than skin, where these fatty acids are found in the secretory product (sebum) of sebaceous glands.  相似文献   

9.
Previously, we have shown that the capping of surface immunoglobulins on murine lymphocytes can be affected by modulating the lipid environment of the surface membrane with free fatty acids. In the present study, murine lymphocytes were depleted of cholesterol by incubation with phospholipid vesicles. As the cellular cholesterol:phospholipid ratio decreased, the capping of the surface immunoglobulin was seen to decrease. This inhibition of capping could not be reversed by calcium and is not accompanied by changes in either the cytoskeletal element alpha-actinin or cellular ATP levels. Incubation of the cholesterol-depleted cells with cholesterol-containing phospholipid vesicles raised both the cholesterol:phospholipid ratio and capping levels to values close to those of untreated control cells. Remarkably, stearic acid, a saturated fatty acid, could also restore the capping levels in the cholesterol-depleted cells. On the basis of the present data and measurements of the fluorescence polarization of the probe diphenyl hexatriene, we propose a model in which the protein(s) involved in capping is located in a gel-like lipid domain, and that removal of cholesterol makes this domain less gel-like and inhibits capping. Restoration of the gel-like nature of this domain by the addition of either cholesterol or stearic acid enables the protein(s) to function normally.  相似文献   

10.
The substrate specificities and selectivities of acyl-CoA synthetasesfrom maturing oilseeds were investigated to reveal fatty acidstructures that the enzymes recognize. The synthetases fromrapeseed (Brassica nap us) and castor bean (Ricinus communis)activated palmitic acid 16:0 most rapidly among the saturatedfatty acids tested. Native unsat-urated fatty acids, oleic 18:1cis-9, linoleic 18:2 cis-9,12 and linolenic acid 18:3 m-9,12,15,were all effectively utilized. Palmitoleic acid 16:1 cis-9 wasalso a good substrate, while myristoleic acid 14:1 cis-9 wasa poor substrate. The activation of erucic acid 22:1 cis-13was very slow. Elaidic acid 18:1 trans-9 was utilized at ratessimilar to those of the cis isomer. The efficiencies of petroselinicacid 18:1 cis-6 were half the efficiencies of oleic acid, whilethe rates of activation of m-vaccenic acid 18:1 cw-11 were comparableto those for oleic acid. These findings suggest that acyl-CoAsynthetases of oilseeds producing long-chain fatty acids strictlyrecognize the molecular structures of fatty acids, i.e., thecarbon-chain length between C16-C18 and the position of thefirst double bond (  相似文献   

11.
The cellular fatty acids of free-living, nitrogen-fixing cyanobacteria belonging to the genera Anabaena and Nostoc were analyzed to differentiate the genera. The fatty acid compositions of 10 Anabaena strains and 10 Nostoc strains that were grown for 12 days on BG-11o medium were determined by gas-liquid chromatography-mass spectroscopy. Of the 53 fatty acids detected, 17 were major components; the average level for each of these 17 fatty acids was at least 0.9% of the total fatty acids (in at least one of the genera). These fatty acids included (with mean percentages in the Anabaena and Nostoc strains, respectively) the saturated fatty acids 16:0 (30.55 and 23.23%) and 18:0 (0.77 and 1.27%); several unsaturated fatty acids, including 14:1 cis-7 (2.50 and 0.11%), 14:1 cis-9 (3.10 and 3.41%), a polyunsaturated 16-carbon (sites undetermined) fatty acid with an equivalent chain length of 15.30 (1.20 and 1.03%), 16:4 cis-4 (0.95 and 0.87%), 16:3 cis-6 (2.16 and 1.51%), 16:1 cis-7 (1.44 and 0.36%), 16:1 cis-9 (6.53 and 18.76%), 16:1 trans-9 (4.02 and 1.35%), 16:1 cis-11 (1.62 and 0.42%), 18:2 cis-9 (10.16 and 12.44%), 18:3 cis-9 (18.19 and 17.25%), 18:1 cis-9 (4.01 and 5.10%), and 18:1 trans-9 (0.92 and 1.94%); and the branched-chain fatty acids iso-16:0 (2.50 and 1.14%) and iso-15:1 (0.34 and 2.05%).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
1. Exogenously supplied, BSA complexed saturated and unsaturated fatty acids were compared for their effects on mitogen-induced DNA synthesis in channel catfish T and B lymphocytes. 2. At "permissive" in vitro temperatures (27 degrees C), high concentrations (greater than or equal to 240 microM) of all the fatty acids used were inhibitory. However, at lower concentrations (80-160 microM), differences were noted in the ability of some fatty acids to modulate mitogen responses. While palmitic acid (16:0) and linoleic acid (18:2) had little effect on LPS-induced B cell- or Con A-induced T cell proliferation, stearic acid (18:0) suppressed while oleic acid (18:1) enhanced T cell responses only. 3. Adding equimolar amounts of 18:0 and 18:1 obviated the effects of singularly added fatty acids on T cell mitogenesis. 4. 18:1 was used to successfully "rescue" approximately 60% of the Con A-induced T cell proliferation normally inhibited at "nonpermissive" in vitro temperatures (17 degrees C). 5. While B cells readily appear to desaturate 18:0 and synthesize unsaturated fatty acids, T cells accumulate comparatively large amounts of 18:0 in membrane associated phospholipids. 6. It is proposed that 18:1 enhances T cell responses at permissive high temperatures and rescues suppressed T cell responses at nonpermissive low temperatures by increasing membrane fluidity.  相似文献   

13.
This investigation was carried out to develop methods for a reverse-phase, high-performance liquid chromatography analysis of the monocarboxylic and dicarboxylic acids produced by permanganate-periodate oxidation of monoenoic fatty acids. Oxidation reactions were performed using [U-14C]oleic acid and [U-14C]oleic acid methyl ester in order to measure reaction yields and product distributions. The 14C-labeled oxidation products consisted of nearly equal amounts of monocarboxylic and dicarboxylic acid (or dicarboxylic acid monomethyl ester), with few side products (yield greater than 98%). Conversion of the carboxylic acids to phenacyl esters proceeded to completion. HPLC of carboxylic acid phenacyl esters was performed using a C18 column with a linear solvent gradient beginning with acetonitrile/water (1/1) and ending with 100% acetonitrile. Excellent resolution was achieved for all components of a mixture of C5 through C12 monocarboxylic acid phenacyl esters and C6 through C11 dicarboxylic acid phenacyl esters. Resolution was also achieved for all components of a mixture of C5 through C12 monocarboxylic acid phenacyl esters and C6 through C11 dicarboxylic acid monomethyl, monophenacyl esters. The resolution obtained by HPLC demonstrates that, for a wide range of monoenoic fatty acids, both products of a permanganate-periodate oxidation can be identified on a single chromatogram. Free fatty acids and fatty acid methyl esters were analyzed with equal success. Neither the oxidation nor the esterification reaction caused detectable hydrolysis of methyl ester. The method is illustrated for free acids and methyl esters of 14:1 (cis-9), 16:1 (cis-9), 18:1 (cis-6), 18:1 (cis-9), and 18:1 (cis-11).  相似文献   

14.
Various unsaturated fatty acids had different effectiveness for maintaining the continued replication of functional mitochondria in an unsaturated fatty acid auxotroph of Saccharomyces cerevisiae (KD115). Certain isomers of octadecenoic acid (i.e., cis-9) and eicosatrienoic acid (i.e.,cis-8,11,14) permitted continued replication of mitochondria and provided cultures that contained only 4 to 5% cells that formed petite colonies. On the other hand, cultures grown with cis-12- or cis-13-octadecenoic acid or cis-11,14,17-eicosatrienoic acid, produced a 12- to 16-fold greater frequency of petite mutants (50-60%) after 8 to 10 generations of growth. The production of the petite mutants occurred despite adequate incorporation of these unsaturated fatty acids into cellular phospholipids and an apparently normal ability to undergo the initial steps in the induction of cellular respiration. The evidence suggests that some cellular processes necessary for continued mitochondrial replication depend on the structural features of the fatty acyl chains as well as the overall content of unsaturated fatty acids in membrane phospholipids. Impairment of that process by certain inadequate fatty acids or by an inadequate supply of a suitable fatty acid leads to a permanent loss of the mitochondrial genome from the cells of subsequent generations.  相似文献   

15.
Transgenic tobacco plants O9 and T16 expressing the yeast acyl-CoA Delta9 desaturase and an insect acyl-CoA Delta11 desaturase, respectively, displayed altered profiles of fatty acids compared to wild-type tobacco plants and marked increases in cis-3-hexenal, a major leaf volatile derived from alpha-linolenic acid (18:3). As expected, O9 and T16 plants had increased levels of the major unsaturated fatty acid products formed by the transgenic desaturases they expressed, viz., palmitoleic acid (16:1(Delta9)) and palmitvaccenic acid (16:1(Delta11)), respectively. In addition, levels of 18:3 lipid declined slightly and the pool of free 18:3, which accounts for about 30% of free fatty acids in wild-type plants, disappeared completely in both transgenics. Both O9 and T16 plants were found to have a two-fold increase in 13-lipoxygenase (13-LOX) activity, which catalyzes the first of two steps leading to hexenal production from 18:3. In O9 and T16 plants, the activity of 9-lipoxygenase and hydroperoxide lyase, the latter catalyzing the formation of cis-3-hexenal from alpha-linolenic acid hydroperoxide, was significantly different from that of the wild-type plants. Although 16:1(Delta9) and 16:1(Delta11) had no direct effects on 13-LOX activity in vitro, cis-3-hexenal production increased in tobacco leaves treated with these fatty acids, suggesting that they may act in vivo by stimulating 13-LOX gene expression.  相似文献   

16.
Frankia isolates from nodules of the genera Casuarina (BR, S21, Thr), Allocasuarina (Allo2), and Gymnostoma (G80) were found to grow exponentially with high biomass yield and minimal sporangia formation in stirred propionate mineral medium when supplemented either with 2.4 μM palmitic acid (C16:0), pentadecanoic (C15:0), heptadecanoic (C17:0), or linoleic (C18:2, cis 9, 12) fatty acids. Strains also grew with lauric (C12:0) or myristic (C14:0) acids, but gave lower biomass yield. Stearic acid (C18:0) produced a good biomass yield, but cultures slowly accumulated sporangia; oleic acid (C18:1, cis-9) was detrimental to growth. Caprylic (C8:0) or capric (C10:0) acids proved to be prejudicial for long-term storage of Frankia strains. In experiments using labeled 1,2-dipalmitoyl phosphatidylcholine and palmitic acid, radioactivity bound rapidly to the insoluble, but solvent-extractable fraction of Frankia cells. In contrast, label from propionic acid accumulated in the cytosolic fraction. Therefore, the beneficial effect of some specific phospatidylcholines or free fatty acids on Frankia growth appears to result from their utilization as building blocks for the membrane, suggesting that membrane biosynthesis may be the limiting step for Frankia growth in unamended propionate mineral medium. Received: 9 October 1995 / Accepted: 24 February 1996  相似文献   

17.
Four samples of freshwater alga Sirodotia (class Rhodophyceae) collected from two distinct streams in the Mahabaleshwar, Satara district (1,732 m a.s.l.) of the Western Ghats of Maharashtra (India) were analysed for their fatty acid content. The presence of 32 fatty acids was revealed, of which 13 were saturated (SFA), 8 were monounsaturated (MUFA) and 11 were polyunsaturated (PUFA) fatty acids. The major finding was the presence of three pharmaceutically and neutraceutically important PUFAs: arachidonic acid (AA), eicosapentanoeic acid (EPA), and docosahexanoiec acid (DHA). The major fatty acids identified were palmitic (16:0), cis-11,14 icodienoic (20:2), behenic (22:0), cis-8,11,14 eicosatrienoic(20:3n6), cis-4,7,10,13,16,19 docosahexanoeic (22:6n3), cis-13,16 docosadienoic (22:2), erucic (22:1n9), -5,8,11,14,17 eicosapentaenoic (20:5n3), trichosonoic (23:0), nervonic (24:0), arachidonic (20:4n6), cis-10 pentadecanoic (15:1), cis-11,14,17 eicosatrienoic (20:3n3), and myristic acid (14:0). The total PUFA contents ranged from 31.45 to 40.37%. The fatty acids were characterised by the relatively high abundance of PUFAs, while C20 unsaturated acids were appreciably more abundant than C18 unsaturated acids. This is the first report on fatty acid profiles of the genus Sirodotia.  相似文献   

18.
cis- and trans-unsaturated fatty acids with 18 carbon atoms (oleic, linoleic, elaidic and linolelaidic acid) inhibited aggregation of washed rabbit platelets stimulated with collagen, arachidonic acid and U46619 when in the same concentration ranges. Thrombin-induced aggregation was not affected by any of them. Saturated fatty acid (stearic acid) had no effect on this response. The inhibition is independent of the induced change in membrane fluidity, since trans-isomers could not induce the change in fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene. Unsaturated fatty acids, except linoleic acid, did not interfere with the formation of thromboxane B2 from exogenously added arachidonic acid. All the unsaturated fatty acids only slightly inhibited the arachidonic acid liberation by phospholipase A2 in platelet lysate. This indicates that the unsaturated fatty acids may block a process after formation of thromboxane A2 in response to collagen and arachidonic acid. The increase in phosphatidic acid formation stimulated with U46619 was inhibited dose dependently by each of the unsaturated fatty acids but that stimulated with thrombin was not affected by any of them. Phospholipase C activity measured by diacylglycerol formation in unstimulated platelet lysate was not inhibited by the fatty acids. The elevation of cytosolic free Ca2+ induced by arachidonic acid or U46619 and Ca2+ influx by collagen were inhibited almost completely at the same concentration as that which inhibited their aggregation. These data suggest that the unsaturated fatty acids were intercalated into the membrane and inhibited collagen- and arachidonic acid-induced platelet aggregation by causing a significant suppression of the thromboxane A2-mediated increase in cytosolic free Ca2+, probably due to interference with the receptor-operated Ca2+ channel.  相似文献   

19.
The degree of plasma membrane fatty acid unsaturation and the copper sensitivity of Saccharomyces cerevisiae are closely correlated. Our objective was to determine whether these effects could be accounted for by differential metal induction of lipid peroxidation. S. cerevisiae S150-2B was enriched with the polyunsaturated fatty acids (PUFAs) linoleate (18:2) and linolenate (18:3) by growth in 18:2- or 18:3-supplemented medium. Potassium efflux and colony count data indicated that sensitivity to both copper (redox active) and cadmium (redox inactive) was increased in 18:2-supplemented cells and particularly in 18:3-supplemented cells. Copper- and cadmium-induced lipid peroxidation was rapid and associated with a decline in plasma membrane lipid order, detected by fluorescence depolarization measurements with the membrane probe trimethylammonium diphenylhexatriene. Levels of thiobarbituric acid-reactive substances (lipid peroxidation products) were up to twofold higher in 18:2-supplemented cells than in unsupplemented cells following metal addition, although this difference was reduced with prolonged incubation up to 3 h. Conjugated-diene levels in metal-exposed cells also increased with both the concentration of copper or cadmium and the degree of cellular fatty acid unsaturation; maximal levels were evident in 18:3-supplemented cells. The results demonstrate heavy metal-induced lipid peroxidation in a microorganism for the first time and indicate that the metal sensitivity of PUFA-enriched S. cerevisiae may be attributable to elevated levels of lipid peroxidation in these cells.  相似文献   

20.
Conjugated linoleic acids (CLAs) reduce fat deposition in several mammalian species. Among the proposed mechanisms for this effect are reduced preadipocyte proliferation and differentiation. We measured proliferation and differentiation of cultured human preadipocytes treated with CLAs. Preadipocytes were differentiated with insulin, hydrocortisone, transferrin, and 10% fetal bovine serum, with isobutyl-methylxanthine included for the first 2 d. The differentiation medium contained 200 microM oleic acid (C18:1), 50 microM cis-9,trans-11-CLA (9,11-CLA), or 50 microM trans-10,cis-12-CLA (10,12-CLA); the negative control medium contained no added fatty acid, and the cells did not differentiate. Cell number increased three to four times during the 17 d of differentiation, but was 30-35% lower in the CLA-treated cells than in the negative control cells. Compared with the negative control cells, differentiation was increased in the cells treated with C18:1 (increased Oil Red O-stained material [OROSM], triacylglycerol, glycerol 3-phosphate dehydrogenase activity [GPDH], peroxisome proliferator-activated receptor-gamma [PPAR gamma] messenger ribonucleic acid [mRNA], and lipoprotein lipase [LPL] mRNA). In effect, the C18:1-treated cells act as a positive control to demonstrate the differentiation capacity of each cell lot. Both 9,11-CLA- and 10,12-CLA-treated cells had increased differentiation (increased OROSM, triacylglycerol, GPDH, PPAR gamma, and LPL) compared with the negative control cells. The data suggest that early in differentiation when de novo fatty acid (FA) synthesis is limited and competition for FAs by membrane and triacylglycerol synthetic pathways is great, human preadipocytes do not differentiate unless a PPAR gamma ligand is added. Either CLA isomer or C18:1 can provide such a ligand.  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