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 46 毫秒
1.
Effects of fatty acids, prostaglandins, and phospholipids on the activity of purified cGMP-stimulated cyclic nucleotide phosphodiesterase from calf liver were investigated. Prostaglandins A2, E1, E2, F1 alpha, and F2 alpha, thromboxane B2, and most phospholipids were without effect; lysophosphatidylcholine was a potent inhibitor. Several saturated fatty acids (carbon chain length 14-24), at concentrations up to 1 mM, had little or no effect on hydrolysis of 0.5 microM [3H]cGMP or 0.5 microM [3H]cAMP with or without 1 microM cGMP. In general, unsaturated fatty acids were inhibitory, except for myristoleic and palmitoleic acids which increased hydrolysis of 0.5 microM [3H]cAMP. The extent of inhibition by cis-isomers correlated with the number of double bonds. Increasing concentrations of palmitoleic acid from 10 to 100 microM increased hydrolysis of [3H]cAMP with maximal activation (60%) at 100 microM; higher concentrations were inhibitory. Palmitoleic acid inhibited cGMP hydrolysis and cGMP-stimulated cAMP hydrolysis with IC50 values of 110 and 75 microM, respectively. Inhibitory effects of palmitoleic acid were completely or partially prevented by equimolar alpha-tocopherol. Palmitelaidic acid, the trans isomer, had only slightly inhibitory effects. The effects of palmitoleic acid (100 microM) were dependent on substrate concentration. Activation was maximal with 1 microM [3H]cAMP and was reduced with increasing substrate; with greater than 10 microM cAMP, palmitoleic had no effect. Inhibition of cGMP hydrolysis was maximal at 2.5 microM cGMP and was reduced with increasing cGMP; at greater than 100 microM cGMP palmitoleic acid increased hydrolysis slightly. Palmitoleic acid did not affect apparent Km or Vmax for cAMP hydrolysis, but increased the apparent Km (from 17 to 60 microM) and Vmax for cGMP hydrolysis with little or no effect on the Hill coefficient for either substrate. These results suggest that certain hydrophobic domains play an important role in modifying the catalytic specificity of the cGMP-stimulated phosphodiesterase for cAMP and cGMP.  相似文献   

2.
The capacity to synthesize both prostaglandins E1 (PGE1) and E2 (PGE2) has been determined in human lung mucoepidermoid carcinoma homogenates when [14C]-fatty acid precursors were added to the incubation medium. Only 10% of the total radioactivity recovered in PGs was found in PGF1 alpha and PGF2 alpha. The experiments were principally focused to inhibit the PGE2 synthesis either with pure eicosapentaenoic (EPA) and docosahexaenoic (DHA) acids or with mixtures of both n-3 fatty acids obtained from fish oil. The results demonstrated that significant inhibitions were found when using 25 microM or a higher concentration of pure EPA or DHA in the incubation medium; however, 5 microM of mixtures of different EPA/DHA ratio caused the same inhibition. The results suggest that EPA and DHA, when added together, may enforce their inhibitory effect on PGE2 synthesis.  相似文献   

3.
A previous study [Berry, M. N., Gregory, R. B., Grivell, A. R. & Wallace, P. G. (1983) Eur. J. Biochem. 131, 215-222] suggested that long-chain fatty acid (palmitate) oxidation by hepatocytes was less sensitive than short-chain fatty acid (hexanoate) oxidation to inhibition by a given concentration of antimycin. Re-examination of this phenomenon showed that palmitate oxidation by hepatocytes could be depressed by antimycin to the same degree as other NAD+-linked substrates, only if the concentration of the inhibitor was raised 2-4-fold. The presence of palmitate also reduced the sensitivity to antimycin of hepatocytes metabolizing lactate or pyruvate. Over the range of fatty acids tested, butyrate (C4) to stearate (C18), only long-chain (greater than C10) fatty acids endowed cells with decreased sensitivity towards antimycin. 2-Bromopalmitate, a non-metabolizable fatty acid, and inhibitor of fatty acid oxidation, also decreased the inhibitory effect of antimycin in cells, suggesting that long-chain fatty acids per se rather than their metabolites, reverse the inhibition by antimycin. Moreover, another inhibitor of fatty acid oxidation, 2-tetradecylglycidic acid, did not diminish the effects of palmitate. Succinate oxidation in isolated mitochondria that had been inhibited by a low concentration of antimycin could be restored by subsequent addition of palmitate or other long-chain fatty acids such as dodecanoate, tetradecanoate and oleate under conditions where fatty acid oxidation was prevented. 2-Bromopalmitate, likewise partially restored antimycin-depressed succinate oxidation. This amelioration of antimycin inhibition was counteracted by the addition of more antimycin and was not seen upon addition of defatted bovine serum albumin, palmitoylcarnitine or octanoate. The total amount of antimycin bound to mitochondria was not affected by the presence of palmitate. The data suggest that long-chain fatty acids are able to interact with the mitochondrial inner membrane in a manner which can relieve the inhibitory effect of antimycin, whether the antimycin is added to the cell or mitochondrial suspension before or after fatty acid addition.  相似文献   

4.
Fatty acid composition of the phospholipids of mouse LM cells grown in suspension culture in serum-free chemically defined medium was modified by supplementing the medium with various fatty acids bound to bovine serum albumin.Following supplementation with saturated fatty acids of longer than 15 carbons (100 μM) profound inhibition of cell growth occurred; this inhibitory effect was completely abolished when unsaturated fatty acids were added at the same concentration. Supplementing with unsaturated fatty acids such as linoleic acid, linolenic acid or arachidonic acid had no effect on the cell growth.Fatty acid composition of membrane phospholipids could be manipulated by addition of different fatty acids. The normal percentage of unsaturated fatty acids in LM cell membrane phospholipids (63%) was reduced to 35–41% following incorporation of saturated fatty acids longer than 15 carbon atoms and increased to 72–82% after addition of unsaturated fatty acids.A good correlation was found between the unsaturated fatty acid content of membrane phospholipids and cell growth. When incorporated saturated fatty acids reduced the percentage of unsaturated fatty acids in membrane phospholipids to less than 50%, severe inhibition of the cell growth was found. Simultaneous addition of an unsaturated fatty acid completely abolished this effect of saturated fatty acids.  相似文献   

5.
beta beta'-Methyl-substituted, C14-C18, alpha, omega-dicarboxylic acids (MEDICA 14-18) were found to inhibit fatty acids and cholesterol synthesis in cultured rat hepatocytes. Maximum inhibition was observed with MEDICA 16, amounting to a 50% decrease in 3H2O and acetate incorporation into fatty acids and cholesterol in the presence of 0.08 mM of the drug added to the culture medium. Inhibition of lipogenesis was not accompanied by inhibition of palmitate or glycerol esterification into neutral lipids and phospholipids. The respective capacities of MEDICA homologues of varying acyl chain length as inhibitors of fatty acid and cholesterol synthesis in cultured rat hepatocytes and in vivo (Bar-Tana, J., Rose-Kahn, G., and Srebnik, M. (1985) J. Biol. Chem. 260, 8404-8410) correlated well with their respective inhibitory effect on liver ATP-citrate lyase. Thus, MEDICA 16 inhibited liver ATP-citrate lyase competitively to citrate with a Ki of 16 microM as compared to a Km of 0.8 mM for the citrate substrate.  相似文献   

6.
The effects of unsaturated fatty acids on the activities of peroxisomal enzymes of Tetrahymena pyriformis were investigated. When saturated fatty acids and the corresponding unsaturated fatty acids (C18) were added to the culture medium at 0.05%, the activities of peroxisomal enzymes [fatty acyl-CoA oxidase (FAO), carnitine acetyltransferase (CAT), isocitrate lyase (ICL), and malate synthase (MS)] were significantly increased. The order of effectiveness was linoleic acid greater than oleic acid greater than stearic acid. However, alpha-linolenic acid and gamma-linolenic acid at the same concentration were lethal to the cells. The inhibitory effect on growth disappeared upon addition of an antioxidant, alpha-tocopherol. Lipid peroxides derived from unsaturated fatty acids induced marked cell lysis. In the presence of a low concentration (0.005%) of linolenic acid the production of lipid peroxide was lower and no inhibitory effect on the growth was observed, while the activities of peroxisomal enzymes participating in lipid metabolism and that of catalase were significantly increased. These results indicate that the peroxisomal enzyme systems related to the beta-oxidations of fatty acids and the glyoxylate cycle are regulated by unsaturated long-chain fatty acids, including linolenic acid, at low concentrations, as well as by saturated fatty acid in the medium.  相似文献   

7.
Treatment of newborn rat calvaria discs with a variety of unsaturated fatty acids led to a 50% enhancement of calcium uptake. Arachidonic acid was effective at lower concentrations than cis-vaccenic or oleic acid, while trans-vaccenic acid and saturated fatty acids did not enhance calcium uptake. Cyclooxygenase inhibitors indomethacin and acetylsalicylic acid abolished the enhancement of calcium uptake seen in response to cis-vaccenic acid and inhibited calcium uptake by otherwise untreated bones. Prostacyclin was found to produce up to 2 fold stimulation of calcium uptake with an EC50 of approximately 0.1 microM. No statistically significant stimulation of calcium uptake was seen in response to PGE2 or PGE1 alpha up to 25 microM, while slight stimulation was produced by 6-keto PGE1 alpha but only at concentrations of 10 microM. Prostacyclin production by calvaria was demonstrated and was stimulated over 50% by cis-vaccenic acid. These results suggest that not only is enhanced prostacyclin production responsible for elevation of calcium uptake in response to unsaturated fatty acids, but also that prostacyclin may be an important regulator of bone calcium homeostasis.  相似文献   

8.
1. Propionate and other unbranched short-chain fatty acids, butyrate, pentanoate, hexanoate and octanoate were found to both stimulate and inhibit active sodium transport by the toad bladder, as measured by the short-circuit current (s.c.c.). 2. Stimulation alone followed addition of low concentrations of fatty acids (0.1-1.0 mM) to either the serosal or mucosal bathing medium; stimulation was also seen after an initial period of inhibition in response to higher concentrations (approx. 5 mM) of some compounds. 3. Inhibition alone followed addition of high concentrations (5-20 mM) of these compounds. The duration and magnitude of the inhibition varied with increasing concentration and chain length of the fatty acid, and was greater following mucosal addition than serosal addition. 4. The inhibitory effect of mucosal propionate increased with decreasing pH of the mucosal bathing medium. 5. Inhibition by the fatty acids was completely reversed upon removing the compound from the bathing medium, and stimulation characteristically followed. 6. In studies designed to evaluate the role of metabolism of the fatty acids in their mucosal inhibitory effects it was found that 14-c-labelled propionate, when added to the mucosal surface of the bladder, was converted to 14-CO2, and mucosal succinate and alpha-oxoglutaric acid at 20 mM inhibited the s.c.c. slightly. However, malonate did not interfere with inhibition by mucosal propionate and two non-metabolizable acids, dimethylpropionate and benzoate, induced inhibition (and no stimulation) of the s.c.c. 7. In the presence of an inhibitory concentration of fatty acid, the ability of the bladder to respond to added pyruvate was reduced in proportion to the reduction in the level of the s.c.c., whereas the natriferic response to vasopressin was largely intact. 8. We conclude that stimulation of sodium transport by propionate and other short-chain fatty acids is due to metabolism of the compounds and provision of energy to the sodium transport mechanism. The basis of the inhibition appears complex. It may in part depend on metabolism of the fatty acids and/or uncoupling of oxidative phosphorylation, with resultant reduction in net ATP production for the sodium transport mechanism. However, the inhibition may also be caused in part by a direct effect on the mucosal entry of sodium into the transporting epithelial cells.  相似文献   

9.
An endogenous inhibitor of the NAD+-dependent 15-hydroxyprostaglandin dehydrogenase was isolated from the 105,000 X g supernatant fraction of lungs of pregnant rabbits following DEAE chromatography. The material was heat stable and was resistant to pronase treatment. The inhibitor contained a mixture of saturated and mono-unsaturated fatty acids and cholesterol with palmitate and oleate representing the major fatty acids in the inhibitory factor. The factor inhibited prostaglandin dehydrogenase activity but had only minor effects on the activity of NAD+-dependent alcohol and lactate dehydrogenases or the NADP+-dependent isocitrate dehydrogenase. In an attempt to develop a greater understanding of the inhibitory action of fatty acids on prostaglandin dehydrogenase activity, a variety of standard fatty acids were examined for their ability to decrease enzymic activity. Oleate and palmitate inhibited enzymic activity by 70% at 10 microM, whereas arachidonate and myristate were only 30% inhibitory at this concentration. A comparison among the 18-carbon-containing fatty acids demonstrated that oleate was more potent than linoleate and linolenate in inhibiting prostaglandin dehydrogenase activity. The coenzyme A derivatives of oleate, linoleate and linolenate were less inhibitory than the free fatty acids.  相似文献   

10.
The possible role of the phospholipase enzymes in the prolactin stimulation of mitogenesis in Nb2 node lymphoma cells was investigated. Two phospholipase inhibitors including quinacrine and alpha-para-dibromoacetophenone (BPB) were employed. Quinacrine at concentrations of 1-5 microM attenuated the magnitude of the PRL stimulation of cell division; at concentrations of 10 microM and above quinacrine abolished the PRL response. BPB at concentrations of 1-10 microM also inhibited the mitogenic effect of PRL in a concentration response fashion. The polyunsaturated fatty acid arachidonic acid partially reversed the inhibitory effects of these drugs. In further studies, exogenously added phospholipase C at concentrations of 5-50 ng/ml was found to potentiate the mitogenic effect of prolactin when prolactin was employed at a concentration that evoked a half-maximal response. By itself, however, phospholipase C had no effect on the rate of cell division. Phospholipase A2 either by itself or in the presence of prolactin was without effect.  相似文献   

11.
The mitoinhibitory effect of fumonisin B1 (FB1) on the mitogenic response of epidermal growth factor (EGF) was investigated in primary hepatocyte cultures with respect to the alterations in the omega6 fatty acid metabolic pathway. Fatty acid analyses of hepatocytes showed that EGF treatment resulted in a significant decrease in the relative levels of 20:4omega6 (arachidonic acid) and an increase in 18:2omega6 (linoleic acid). Supplementation of the hepatocyte cultures with 20:4omega6 in the absence of EGF resulted in an increase in the total omega6 and omega6/omega3 fatty acid ratio. Addition of 20:5omega3 (eicosapentaenoic acid) resulted in an increase of the relative levels of the long chain omega3 fatty acids at the expense of the omega6 fatty acids. When 20:4omega6 and 20:5omega3 was added in the presence of EGF, the mitogenic response of EGF was increased and decreased respectively. When compared to the fatty acid profiles in the absence of EGF, the decreased mitogenic response coincided with a decrease of total omega6 fatty acids and total polyunsaturated fatty acids (PUFA). In addition, the saturated and mono-unsaturated fatty acids increased and the polyunsaturated/saturated (P/S) fatty acid ratio decreased which implied a more rigid membrane structure. Addition of prostaglandin E2 (PGE2) and prostaglandin E1 (PGE1) stimulated and inhibited the mitogenic response respectively. Ibuprofen, a known cyclooxygenase inhibitor, and FB1 inhibited the EGF-induced mitogenic response in a dose-dependent manner. The mitoinhibitory effect of FB1 on the EGF response was counteracted by the addition of PGE2. FB1 also disrupts the omega6 fatty acid metabolic pathway in primary hepatocytes, resulting in the accumulation of C18:2omega6 in phospatidylcholine and triacylglicerol. The disruption of the omega6 fatty acid metabolic pathway and/or prostaglandin synthesis is likely to be an important event in the mitoinhibitory effect of FB1 on growth factor responses.  相似文献   

12.
Murine thioglycollate-elicited peritoneal macrophages were cultured in the presence of a variety of fatty acids added as complexes with bovine serum albumin. All fatty acids tested were taken up readily by the cells and both neutral and phospholipid fractions were enriched with the fatty acid provided in the medium. This generated a range of cells enriched in saturated, monounsaturated or polyunsaturated fatty acids, including n-3 acids of fish oil origin. Saturated fatty acid enrichment enhanced macrophage adhesion to both tissue culture plastic and bacterial plastic compared with enrichment with polyunsaturated fatty acids. Macrophages enriched with the saturated fatty acids myristate or palmitate showed decreases of 28% and 21% respectively in their ability to phagocytose unopsonized zymosan particles. Those enriched with polyunsaturated fatty acids showed 25-55% enhancement of phagocytic capacity. The greatest rate of uptake was with arachidonate-enriched cells. Phagocytic rate was highly correlated with the saturated/unsaturated fatty acid ratio, percentage of polyunsaturated fatty acid and index of unsaturation, except for macrophages enriched with fish-oil-derived fatty acids; they showed lower phagocytic activity than expected on the basis of their degree of unsaturation. These results suggest that membrane fluidity is important in determining macrophage adhesion and phagocytic activity. However, in the case of phagocytosis, this effect may be partially overcome if the cells are enriched with fish-oil-derived fatty acids. Thus it may be possible to modulate the activity of cells of the immune system, and so an immune response, by dietary lipid manipulation.  相似文献   

13.
We examined the influence of prostaglandins on the initiation of proliferation of growth-arrested human adult aortic and fetal smooth muscle cells. Prostaglandins of the E series (25 nM) exerted a significant (p less than or equal to 0.05) inhibitory effect on DNA synthesis. Inhibition was observed when PGE1 was added in the G1 phase of the cell cycle. PGE1 had no effect when added once DNA synthesis had started. Thus prostaglandins of the E series may inhibit the responsiveness of smooth muscle cells to the mitogenic action of critical growth factors, such as PGDF. This inhibitory response is cell-cycle dependent. Once smooth muscle cells have entered S phase, PGE1 is no longer effective. Our data also suggest that cAMP is involved in the PGE1-induced growth inhibition, since concomitant with PGE1 addition, cAMP levels rose rapidly; addition of the cAMP analogue db-cAMP resulted in a cell-cycle-dependent inhibition pattern comparable to that observed with PGE1.  相似文献   

14.
Summary The protective influence of bovine serum albumin against growth inhibition caused by fatty acids was studied in human hepatoma (HepG2) and immortalized human kidney epithelial (IHKE) cells. In general, growth inhibition by unsaturated fatty acids (0.15 mmol/liter) increased with increasing number of double bonds. For HepG2 cells crude albumin (1g/100 ml) did not greatly modify growth inhibition by arachidonic, eicosapentaenoic, and docosahexaenoic acid. With oleic, linoleic, and linolenic acids, crude and defatted albumin stimulated cell growth. In contrast, for IHKE cells both albumins counteracted growth inhibition by unsaturated fatty acids to approximately the same extent. When HepG2 cells were cultured in the presence of saturated fatty acids (0.3 mmol/liter), C2, C6, and C8 had no or little inhibitory effect. C10 and C12 inhibited cell growth appreciably, whereas C14, and especially C16, had poor inhibitory effects. Crude albumin counteracted growth inhibition by all these fatty acids. In contrast, defatted albumin had little or no effect (except against C10 and C12), and even increased the growth inhibition by C14 and C16. With unsaturated fatty acids there seemed to be an inverse relationship between cell growth and the concentration of thiobarbituric acid reactive substances (TBARS) in media. Vitamin E abolished growth inhibition (and the increase in TBARS concentration) by unsaturated fatty acids. The complex interaction between fatty acids and albumins calls for great caution when interpreting data on growth effects.  相似文献   

15.
Murine fibroblasts, LM cells, were cultured in suspension with laurate (12:0), myristate (14:0), palmitate (16:0), palmitoleate (16:1), or palmitate + palmitoleate (16:0 + 16:1) bound to fatty acid-free bovine serum albumin. Supplementation with saturated fatty acids decreased the ratio of unsaturated/saturated fatty acids in membrane phospholipids as much as 3.4-fold (palmitate-enriched cells). Concomitantly fluorescence polarization, absorption-corrected fluorescence, and relative fluorescence efficiency of the fluorescence probe molecule, β-parinaric acid, increased 1.5-, 2.9-, and 1.8-fold, respectively, in the membrane phospholipids. Unsaturated fatty acid (palmitoleate) increased the unsaturated/saturated fatty acid ratio by 20% but did not significantly alter the fluorescence parameters. When the cells were fed mixtures of palmitate and palmitoleate, the unsaturated/saturated fatty acid ratio of the membrane phospholipids and the above fluorescence parameters had values intermediate between those if each fatty acid had been fed separately. All fatty acid supplements caused a loss of two characteristic temperatures in Arrhenius plots of relative fluorescence efficiency. However, no shifts or appearance of new characteristic temperatures occurred. The break points at approximately 42, 37, and 22 °C were essentially un-altered. The data were consistent with the possibility that LM cells were unable to maintain constant fluidity, as indicated by fluorescence polarization, when supplemented with different fatty acids. A good correlation could be made between the phospholipid unsaturated/ saturated fatty ratio, the fluorescence polarization, and the toxicity elicited by different fatty acid supplements.  相似文献   

16.
The activities of rat brain prostaglandin D synthetase and swine brain prostaglandin D2 dehydrogenase were inhibited by some saturated and unsaturated fatty acids. Myristic acid was most potent among saturated straight-chain fatty acids so far tested. The IC50 values of this acid were 80 microM for prostaglandin D synthetase and 7 microM for prostaglandin D2 dehydrogenase, respectively. Little inhibition was found with methyl myristate and myristyl alcohol. The IC50 values of these derivatives were more than 200 microM for both enzymes, suggesting that the free carboxyl group was essential for the inhibition. The effects of cis double bond structure of fatty acids on the inhibition potency were examined by the use of the carbon 18 and 20 fatty acids. The inhibition potencies for both enzymes increased with the number of cis double bonds; the IC50 values of stearic, oleic, linoleic and linolenic acid were, respectively, more than 200, 60, 30 and 30 microM for prostaglandin D synthetase, and 20, 10, 8.5 and 7 microM for prostaglandin D2 dehydrogenase. Arachidonic acid also inhibited the activities of both enzymes with respective IC50 values of 40 microM for prostaglandin D synthetase and 3.9 microM for prostaglandin D2 dehydrogenase, while arachidic acid showed little inhibition. The kinetic studies with myristic acid and arachidonic acid demonstrated that the inhibition by these fatty acids was competitive and reversible for both enzymes. Myristic acid and other fatty acids also inhibited the activities of several enzymes in prostaglandin metabolism, although to a lesser extent. The IC50 values of myristic acid for prostaglandin E isomerase, thromboxane synthetase and NAD-linked prostaglandin dehydrogenase (type I) were 200, 700 and 100 microM, respectively. However, this fatty acid showed little inhibition on fatty acid cyclooxygenase (20% at 800 microM), glutathione-requiring prostaglandin D synthetase from rat spleen (20% at 800 microM), and NADP-linked prostaglandin dehydrogenase (type II) (no inhibition at 200 microM).  相似文献   

17.
Long chain fatty acids were found to inhibit (Na+ + K+)-ATPase prepared from rat heart. Unsaturated and polyunsaturated fatty acids were more inhibitory than saturated fatty acids with myristic acid being the most inhibitory saturated fatty acid tested and linoleic the most inhibitory unsaturated fatty acid. As an example of fatty acid modification of the enzyme, inhibition of (Na+ + K+)-ATPase by oleate was examined. When compared to ouabain, inhibition of (Na+ + K+)-ATPase by oleate was found to be similar in that both were dependent on K+ concentration, but, in contrast to the almost instantaneous inhibition by ouabain, oleate inhibition was a slow process requiring over 20 min incubation at 37 degrees to produce maximum inhibition. Inhibition of rat heart (Na+ + K+)-ATPase by oleate was found to be readily reversible by washout. In the presence of albumin an oleate/albumin molar ratio greater than 7.5 was required for inhibition to occur. The activity of rat heart (Na+ + K+)-ATPase had a temperature optimum above 40 degrees and a discontinuous Arrhenius' plot with a transition temperature of 25 degrees. In the presence of oleate, however, the enzyme's optimum temperature decreased to below 40 degrees, the activation energy of the reaction at temperatures below 25 degrees was lowered from 24.7 kcal/mol to 12.6 kcal/mol and the enzyme had a linear Arrhenius' plot. The possibility of in vivo inhibition of cardiac (Na+ + K+)-ATPase under conditions of elevated fatty acids is discussed.  相似文献   

18.
When phospholipases of plasma membranes are activated by certain stimuli, unsaturated fatty acids are liberated. Because unsaturated fatty acids enhance the transmembrane movement of calcium ions, the fatty acids released may modulate intracellular calcium homeostasis in various cells, including neutrophils. To determine the physiological function of these unsaturated fatty acids, we studied the effects of various fatty acids on superoxide generation and on changes in intracellular calcium contents of guinea pig neutrophils. Some unsaturated fatty acids, arachidonate and linoleate, stimulated the rate of superoxide generation concomitant with the increase in the amount of intracellular calcium. In contrast, the saturated fatty acid, myristate, stimulated the generation of superoxide without affecting the content of intracellular calcium. The stimulating actions of arachidonate and myristate were increased dramatically by the presence of a low concentration (1 microM) of extracellular calcium ion. The rate of superoxide generation in fatty acid-treated neutrophils was inhibited by chlorpromazine, an inhibitor of such calcium-binding proteins as C-kinase. These and other observations suggest that liberated unsaturated fatty acids increase the amount of intracellular calcium and enhance C-kinase activity also that the increased activity of the enzyme is involved in the chain of events leading to the stimulation of superoxide generation in fatty acid-treated neutrophils.  相似文献   

19.
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   

20.
Summary Sulfo-N-succinimidyl derivatives of the long-chain fatty acids, oleic and myristic, were synthesized and covalently reacted with isolated rat adipocytes. The plasma membrane proteins labeled by these compounds and the effect of labeling on the transport of long-chain fatty acids were investigated. Sulfo-N-succinimidyl oleate (SSO) and myristate (SSM) inhibited the transport of fatty acids (by about 70%). Inhibition of fatty acid transport was not a result of alterations in cell integrity, as intracellular water volume was not changed. It did not reflect effects on fatty acid metabolism, since it was observed under conditions where greater than 90% of the fatty acid taken up was recovered in the free form. The inhibitory effect was specific to the fatty acid transport system, as the transport of glucose and the permeation of retinoic acid, a substance with structural similarities to long-chain fatty acids, were unaffected. Sulfosuccinimidyl oleate reacted exclusively with a plasma membrane protein with an apparent size of 85 kDa while sulfosuccinimidyl myristate also labeled a 75-kDa while sulfosuccinimidyl myristate also labeled a 75-kDa protein. These proteins were among the ones labeled by diisothiocyanodisulfonic acid (DIDS) which also inhibits fatty acid transport irreversibly. The data suggest that the 85-kDa protein, which is the only one labeled by all three inhibitors is involved in facilitating membrane permeation of long-chain fatty acids.  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