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 role of arachidonic acid in rat heart cell metabolism   总被引:4,自引:0,他引:4  
Although it is known that arachidonic acid accumulates in the ischemic myocardium and that cardiac prostaglandin formation from the precursor arachidonic acid is altered during disease states, the role of arachidonic acid in the myocyte itself is not yet clear. Using isolated Ca-tolerant adult rat heart muscle cells, we were able to study cardiac metabolism of arachidonic acid without the effects induced by endothelial or other non-muscle tissue. Myocytes rapidly incorporate arachidonic acid as well as other fatty acids into their lipid pools, the predominant acceptor being the triacylglycerols at an extracellular fatty acid concentration of 20 microM. As exogenous arachidonic acid is decreased, the distribution pattern shifts to favor phospholipid esterification. Cardiocyte prostaglandin production from arachidonic acid added to the incubation medium was limited (less than 1% conversion of added arachidonic acid) and lipoxygenase pathway activity was not detected. Oxidation rates of arachidonic acid were 3-fold lower than for palmitic acid, indicating that it is of secondary importance in energy-yielding reactions. Our results suggest that arachidonic acid serves primarily as a structural component of myocardial membranes and that its release during ischemia would permit its use as a substrate for prostaglandin production by coronary vascular tissue.  相似文献   

2.
B. Janistyn 《Planta》1982,155(4):342-344
Gas chromatographic-mass spectroscopic evidence is presented for the presence of arachidonic acid ((all-Z)-5,8,11,14-eicosatetraenoic acid) in virgin wheat-germ oil. The amount of arachidonic acid in wheat-germ oil, determined as methyl ester using gas chromatographic detection, was quantified with reference to a standard methyl ester from authentic arachidonic acid. The presence of arachidonic acid and prostaglandins in higher plants is discussed.  相似文献   

3.
Whereas antibipolar drug administration to rats reduces brain arachidonic acid turnover, excessive N-methyl-d-aspartate (NMDA) signaling is thought to contribute to bipolar disorder symptoms and may increase arachidonic acid turnover in rat brain phospholipids. To determine whether chronic NMDA would increase brain arachidonic acid turnover, rats were daily administered NMDA (25 mg/kg, ip) or vehicle for 21 days. In unanesthetized rats, on day 21, [1-(14)C]arachidonic acid was infused intravenously and arterial blood plasma was sampled until the animal was euthanized at 5 min and its microwaved brain was subjected to chemical and radiotracer analysis. Using equations from our in vivo fatty acid model, we found that compared with controls, chronic NMDA increased the net rate of incorporation of plasma unesterified arachidonic acid into brain phospholipids (25-34%) as well as the turnover of arachidonic acid within brain phospholipids (35-58%). These changes were absent at 3 h after a single NMDA injection. The changes, opposite to those after chronic administration of antimanic drugs to rats, suggest that excessive NMDA signaling via arachidonic acid may be a model of upregulated arachidonic acid turnover in brain phospholipids.  相似文献   

4.
The inhibitory mechanism of high levels of exogenously added arachidonic acid on activation of washed human platelets was investigated. While low levels of arachidonic acid (5-10 microM) induced aggregation, ATP secretion and increase in cytoplasmic free Ca2+ concentration (first phase of activation), these platelet responses did not occur significantly at high concentrations (30-50 microM). However, much higher concentrations than 80 microM again elicited these responses (second phase). The first phase of platelet activation was inhibited by cyclooxygenase inhibitor, indomethacin, whereas the second one was independent of such treatment. Thromboxane B2 was produced dose-dependently until reaching a plateau at arachidonic acid concentrations higher than 20 microM, irrespective of the lack of aggregation and secretion at high concentrations. After that the amount of free arachidonic acid which remained unmetabolized in platelets gradually increased. High concentrations of arachidonic acid as well as other polyunsaturated fatty acids caused desensitization of platelets in response to U46619, and also depressed the specific [3H]U46619-binding to the receptor as well as other polyunsaturated fatty acids. The amount free arachidonic acid needed in platelets to suppress [3H]U46619 binding corresponded to that needed to inhibit platelet aggregation. Furthermore, arachidonic acid dose-dependently induced fluidization of lipid phase of platelet membranes as detected by 1,6-diphenyl-1,3,5-hexatriene. These results suggest that the inhibition of platelet response by high levels of arachidonic acid can be attributed to interference with endoperoxide/thromboxane A2 binding to the receptor, probably due to perturbation of the membrane lipid phase due to excess amounts of free arachidonic acid remaining in the membranes.  相似文献   

5.
The polyunsaturated fatty acid composition of Brugia malayi microfilariae was analyzed by gas chromatography and compared to that of sera from B. malayi-infected jirds. The essential fatty acid, linoleic acid (18:2 omega 6), was the most abundant fatty acid present in both microfilarial total lipids and phospholipids as well as in jird sera. In contrast, arachidonic acid (20:4 omega 6), as well as the 18:3 omega 6, 20:2 omega 6, and 20:3 omega 6 intermediates that are formed in the enzymatic conversion of linoleic acid to arachidonic acid, were proportionally more abundant in microfilariae than in jird sera. To assess the capacity of microfilariae to transform linoleic acid into arachidonic acid, B. malayi microfilariae were incubated with [14C]linoleic acid. Microfilarial lipids were extracted and resolved by high-pressure liquid chromatography and thin-layer chromatography. A portion of [14C]linoleic acid incorporated by microfilariae was converted to [14C]arachidonic acid. Thus, microfilariae can not only incorporate exogenous arachidonic acid, as previously demonstrated, but can also synthesize arachidonic acid from exogenous linoleic acid. The capacity of microfilariae to utilize specific host polyunsaturated fatty acids raises the possibility that intravascular filarial parasites may synthesize eicosanoid metabolites of arachidonic acid that could mediate filarial-host cell interactions.  相似文献   

6.
T Daum  M S Rohrbach 《FEBS letters》1992,309(2):119-122
Zymosan, which is composed primarily of alpha-mannan and beta-glucan polymers, is a well recognized activator of macrophages. The type receptor by which unopsonized zymosan induces arachidonic acid release was investigated. It was found that particulate beta-glucan and zymosan stimulated an identical dose-dependent release of arachidonic acid. This release of arachidonic acid by zymosan was blocked by soluble beta-glucans whereas soluble mannan had no effect. This inhibition was not due to a general toxic effect of the soluble beta-glucans as they had no effect on calcium ionophore-induced release of arachidonic acid. Beta-glucan-induced fatty acid release from these cells was shown to be fairly specific for arachidonic acid. These data reveal that zymosan stimulates the specific release of arachidonic acid from rabbit alveolar macrophages, at least in part, via a beta-glucan receptor.  相似文献   

7.
Studies from our laboratory demonstrated that the free radical scavenger, nitro blue tetrazolium, and iron chelators, such as dypyrydil, are potent inhibitors of arachidonic acid oxidation and platelet function. In the present study, we have evaluated the effects of known antioxidants, such as butylated hydroxyanisol (BHA), butylated hydroxytoluene (BHT), and diphenylamine, on arachidonic acid metabolism and platelet function. Diphenylamine, a common dye intermediate used in hair color formulations, was the most potent inhibitor of arachidonic acid metabolism by platelet cyclooxygenases. Diphenyl and BHA were also potent inhibitors of arachidonic acid oxidation. Other diphenyl analogues and BHT were relatively poor inhibitors of arachidonic-mediated platelet activation. Results of this study, as well as those of our earlier studies, suggest that antioxidants and iron chelators prevent arachidonic acid metabolism and alter platelet function by interfering with the heme/arachidonic acid interaction and blocking cyclooxygenase metabolites essential for the formation of thromboxane A2, a potent platelet agonist.  相似文献   

8.
Because products of arachidonic acid metabolism, particularly the PG, have been implicated as modulators of growth and differentiation of adult thymocytes, we investigated relations between metabolism of arachidonic acid and growth, as well as differentiation, of thymocytes during fetal thymic organ culture. Fetal thymic cells synthesized immunoreactive PGE2 during organ culture and were found to be capable of metabolizing exogenous arachidonic acid to products that cochromatographed with authentic 6-keto-PGF1 alpha, PGE2, PGF2 alpha. Synthesis of these products and growth and expression of Thy-1 and Lyt-1 Ag were inhibited by culture of fetal thymic lobes with indomethacin, a cyclooxygenase inhibitor, as well as meclofenamate and eicosatetraynoic acid, inhibitors of cyclooxygenase and lipoxygenase pathways of arachidonic acid metabolism. Only indomethacin inhibited expression of Lyt-2. Culture with eicosatetraynoic acid also inhibited the capacity of thymic lobes to synthesize 15-hydroxyeicosatetraenoic acid-like products. The inhibitory effects of indomethacin on growth and expression of Thy-1 were partially reversed by simultaneous addition of arachidonic acid. Thus, fetal thymic cells appear to require an intact cyclooxygenase, and possibly lipoxygenase, pathway of arachidonic acid metabolism for growth and differentiation. These data also provide evidence that Lyt-1 and Lyt-2 may be regulated by different requirements with respect to arachidonic acid metabolism.  相似文献   

9.
Rat hepatocytes in primary culture were incubated with a mixture of linoleic and arachidonic acid at various total fatty acid/serum albumin molar ratios. Mixed fatty acids were taken up at the same rate and distributed with the same pattern as fatty acids added separately. The rates of total uptake, incorporation into hepatocyte and secreted triacylglycerols and beta-oxidation were linearly related to the fatty acid/albumin ratios, whereas the rate of incorporation into phospholipids was saturable. Neither the uptake rate nor the distribution of both fatty acids considered together varied with the arachidonic acid/linoleic acid molar ratio. Changes in this ratio and in the uptake rate led to significant variations in the respective fate of the fatty acids. The preferential channelling of arachidonic acid versus linoleic acid into beta-oxidation and phosphatidylinositol was greatest at a low uptake rate and then decreased as the uptake rose. Conversely, the preferential channelling of arachidonic acid versus linoleic acid into phosphatidylcholine, but not phosphatidylethanolamine, increased with the uptake rate. Moreover, both arachidonic acid and linoleic acid were preferentially incorporated into the 1-palmitoyl molecular species of phosphatidylcholine and phosphatidylethanolamine at a low uptake rate, and of phosphatidylcholine at a high uptake rate. This could be related to the synthesis of biliary phosphatidylcholine, of which 1-palmitoyl-2-linoleoyl and 1-palmitoyl-2-arachidonoyl are the main molecular species. Linoleic and arachidonic acid were selectively distributed into distinct metabolic pools of triacylglycerol, the intrahepatocyte pool which preferentially incorporated linoleic acid at a low uptake rate and the secreted pool in which the relative enrichment of arachidonic acid increased with the uptake rate. This strengthens the central role of hepatic secretion in the supply of arachidonic acid to peripheral tissues.  相似文献   

10.
Stimulation of rabbit polymorphonuclear leucocytes with A23187 causes phospholipase C mediated breakdown of polyphosphoinositides, as evidenced by accumulation of [3H]inositol-labelled inositol bisphosphate and inositol trisphosphate. At the same time the polyphosphoinositides and the products of their breakdown, diacylglycerol and phosphatidic acid, label rapidly with radioactive arachidonic acid. Enhancement of polyphosphoinositide labelling is not as great as enhancement of diacylglycerol or phosphatidic acid labelling, suggesting additional early activation of a second independent synthetic pathway to the last named lipids. Experiments using double (3H/14C) labelling, to distinguish pools with different rates of turnover, suggest the major pool of arachidonic acid used for synthesis of lipoxygenase metabolites turns over more slowly than arachidonic acid in diacylglycerol, but at about the same rate as arachidonic acid esterified in phosphatidylcholine or phosphatidylinositol. Further, when cells are prelabelled with [14C]arachidonic acid, then stimulated for 5 min, it is only from phosphatidylcholine, and to a lesser extent phosphatidylinositol, that radiolabel is lost. Release of arachidonic acid is probably via phospholipase A2, since it is blocked by the phospholipase A2 inhibitor manoalide. The absence of accumulated lysophosphatides can be explained by reacylation and, in the case of lysophosphatidylinositol, deacylation. The importance of phospholipase A2 in phosphatidylinositol breakdown contrasts with the major role of phospholipase C in polyphosphoinositide hydrolysis. Measurements of absolute free fatty acid levels, as well as studies showing a correlation between production of radiolabelled hydroxyeicosatetraenoic acids and release of radiolabel from the phospholipid pool, both suggest that hydrolysis of arachidonic acid esterified into phospholipids is the limiting factor regulating formation of lipoxygenase metabolites. By contrast with A23187, fMet-Leu-Phe (a widely used polymorphonuclear leucocyte activator) is a poor stimulant for arachidonic acid release unless a 'second signal' (e.g. cytochalasin B, or a product of A23187-stimulated cells) is also present. In the presence of cytochalasin B, fMet-Leu-Phe, like A23187, stimulates release of radiolabelled arachidonic acid principally from phosphatidylcholine.  相似文献   

11.
A gas-liquid chromatography-flame ionization method is described for measuring arachidonic acid in plasma using dihomo-gamma-linolenic acid as an internal standard. We found this technique to -e reproducible, and quicker and superior to previously described techniques because of the similar physico-chemical properties of the unsaturated fatty acid internal standard and arachidonic acid. In addition, we observed that the use of the saturated fatty acid, n-tricosanoic acid, was unsatisfactory as an internal standard because of its poor extractability from plasma as compared to arachidonic acid.  相似文献   

12.
The relative contributions of arachidonic acid and protein kinase C during GnRH-stimulated LH release were investigated in cultured rat anterior pituitary cells. Maximal or near-maximal concentrations of arachidonic acid or the phorbol ester, 12-O-tetradecanoylphorbol 13-acetate, were less effective than a maximal dose of GnRH in stimulating LH release. However, the effect of a combination of arachidonic acid and phorbol ester was equivalent with that of GnRH. The protein kinase C inhibitor, retinal, significantly reduced GnRH- and phorbol-induced, but not arachidonic acid-stimulated, LH release. The lipoxygenase inhibitors, 5,8,11,14-eicosatetraynoic acid and nordihydroguaiaretic acid, partially inhibited GnRH- and arachidonic acid-stimulated, but not phorbol-induced, LH secretion. Simultaneous addition of retinal and either lipoxygenase inhibitor completely abolished LH responses elicited by GnRH, as well as by combined treatment with arachidonic acid and the phorbol ester. These results suggest that hormone release is mediated by phospholipid-dependent mechanisms that are coordinated during the stimulation of LH secretion by GnRH.  相似文献   

13.
花生四烯酸作为一种重要的多价不饱和脂肪酸,因其具有多种生理功能而被认为是潜在的食品添加剂和药物。近年来,利用高山被孢霉合成花生四烯酸已成为研究热点。前期相关研究主要集中在菌种选育及发酵调控方面。随着研究的不断深入,关于高山被孢霉合成花生四烯酸的代谢途径的研究取得了较大进展。以下简要概述前期工作进展,着重论述花生四烯酸合成途径的关键酶及其高山被孢霉的遗传改造的研究情况,包括生物合成花生四烯酸代谢途径、关键酶及其应用、高山被孢霉的遗传操作系统的构建以及遗传改造的应用,并对其研究前景进行了展望。  相似文献   

14.
The migration of human T lymphocytes, assessed in modified Boyden chambers, was chemokinetically stimulated by arachidonic acid in a dose-related manner that achieved a peak level of 127 ± 34% enhancement (mean ± SD) at 8 μM arachidonic acid. The chemokinetic effect was dependent on the metabolism of the arachidonic acid by the T lymphocytes as derivatives of arachidonic acid that do not serve as prostaglandin and thromboxane precursors were without effect, while the cyclo-oxygenase inhibitors indomethacin (ID50 = 10 μM) and 5,8,11,14-eicosatetraynoic acid (ETYA) (ID50 = 20 μM) suppressed the stimulation of migration by arachidonic acid. The cyclo-oxygenase product 12-l-hydroxy-5,8,10-heptadecatrienoic acid (HHT) reproduced part of the chemokinetic effect of arachidonic acid, but the lipoxygenase product 12-l-hydroxy-5,8,10,14-eicosatetraenoic acid (HETE) as well as PGE2, PGF, and thromboxane B2 had no stimulatory activity. The ability of ETYA, but not indomethacin, to suppress the migration of unstimulated T lymphocytes suggested that a lipoxygenase metabolite of endogenous arachidonic acid contributes to the maintenance of their normal levels of spontaneous migration.  相似文献   

15.
Prostaglandin synthesis in endothelial cells may be initiated by the addition of exogenous substrate (arachidonic acid) or by addition of thrombin or the CA2+-ionophore A23187, which leads to prostacyclin formation from endogenous substrates. We noticed that endothelial cells produce more than twice the amount of prostacyclin when incubated with thrombin and arachidonic acid together than with arachidonic acid alone. In addition, it was found that the thrombin-induced conversion of endogenous substrates was inhibited by exogenous arachidonic acid. This means that the conversion of exogenous added arachidonic acid to prostacyclin was stimulated by thrombin. This activation of the enzymes involved in prostacyclin synthesis lasted about 5 min and could be inhibited by phospholipase inhibitors such as mepacrine and p-bromophenyl-acylbromide but not by the cAMP analogue dibutyryl cAMP, an inhibitor of arachidonic acid release from cellular phospholipids. These data demonstrate that, in addition to causing release of endogenous substrate, thrombin and the Ca2+-ionophore also activate the enzyme system involved in the further transformation of arachidonic acid.  相似文献   

16.
Purified cytochromes P-450 (LM2 and PB-B2) in a reconstituted system and epoxide hydrolase were recently found to metabolize arachidonic (eicosatetraenoic) acid to four vicinal dihydroxyeicosatrienoic acids. These metabolites were chemically synthetized from octadeuterated arachidonic acid and employed as internal standards for mass fragmentography. Isolated rat hepatocytes and renal cells were incubated with arachidonic acid (0.1 mM; 37°C, 15 min) and, following extractive isolation and reversed-phase HPLC, formation of 11,12-dihydroxy-5,8,14-eicosatrienoic acid and 14,15-dihydroxy-5,8,11-eicosatrienoic acid was demonstrated by mass fragmentography using a capillary GC column. Furthermore, these diols were also detected in rabbit liver and renal cortex and they therefore appear to be formed endogenously. Formation of vicinal diols was also studied in cell free systems. Rabbit liver and renal cortical microsomes were incubated with NADPH (1 mM) and arachidonic acid (0.15 mM) for 15 min at 37°C and, besides 11,12-dihydroxy- and 14,15-dihydroxyeicosatrienoic acid, small amounts of 8,9-dihydroxy- and 5,6-dihydroxyeicosatrienoic acid could be detected by mass fragmentography. Renal as well as hepatic monooxygenases can thus epoxidize each of the four double bonds of arachidonic acid. In contrast, rabbit lung microsomes and NADPH metabolize arachidonic acid mainly to prostaglandins and 19-hydroxy- and 20-hydroxyarachidonic acid, while only small amounts of 11,12-dihydroxyeicosatrienoic acid could be found. Monooxygenase metabolism of arachidonic acid by epoxidation might therefore be a significant pathway for the metabolism of this essential fatty acid in isolated rat renal cells and hepatocytes but presumably not in the lung.  相似文献   

17.
Studies from our laboratory have suggested a role for ferrous iron in the metabolism of arachidonic acid and demonstrated that inhibitors of prostaglandin synthesis exert their effect by complexing with the heme group of cyclooxygenase. Docosahexaenoic acid (DHA) is a potent competitive inhibitor of arachidonic acid metabolism by sheep vesicular gland prostaglandin synthetase. In this study we have evaluated the effect of exogenously added DHA on platelet function and arachidonic acid metabolism. DHA at 150 microM concentration inhibited aggregation of platelets to 450 microM arachidonic acid. At this concentration DHA also inhibited the second wave of the platelet response to the action of agonists such as epinephrine, adenosine diphosphate and thrombin. Inhibition induced by this fatty acid could be overcome by the agonists at higher concentrations. DHA inhibited the conversion of labeled arachidonic acid to thromboxane by intact, washed platelet suspensions. However, platelets in plasma incubated first with DHA then washed and stirred with labeled arachidonate generated as much thromboxane as control platelets. These results suggest that the polyenoic acids, if released in sufficient quantities in the vicinity of cyclooxygenase, could effectively compete for the heme site and inhibit the conversion of arachidonic acid.  相似文献   

18.
We have investigated the direct effect of arachidonic acid on cholesterol transport in intact cells or isolated mitochondria from steroidogenic cells and the effect of cyclic-AMP on the specific release of this fatty acid inside the mitochondria. We show for the first time that cyclic-AMP can regulate the release of arachidonic acid in a specialized compartment of MA-10 Leydig cells, e.g. the mitochondria, and that the fatty acid induces cholesterol transport through a mechanism different from the classical pathway. Arachidonic acid and arachidonoyl-CoA can stimulate cholesterol transport in isolated mitochondria from nonstimulated cells. The effect of arachidonoyl-CoA is inhibited by the reduction in the expression or in the activity of a mitochondrial thioesterase that uses arachidonoyl-CoA as a substrate to release arachidonic acid. cAMP-induced arachidonic acid accumulation into the mitochondria is also reduced when the mitochondrial thioesterase activity or expression is blocked. This new feature in the regulation of cholesterol transport by arachidonic acid and the release of arachidonic acid in specialized compartment of the cells could offer novel means for understanding the regulation of steroid synthesis but also would be important in other situations such as neuropathological disorders or oncology disorders, where cholesterol transport plays an important role.  相似文献   

19.
Macrophage microsomes catalyzed the transfer of arachidonic acid (20:4) from 1,2-diacyl-glycerophosphocholine (GPC) to 1-alkyl-GPC (lyso platelet-activating factor). This enzyme reaction did not require the presence of cofactors such as Co A. Free arachidonic acid or linoleic acid-labeled phospholipids failed to act as the acyl donor. These results suggest that the reaction is a CoA-independent direct transfer of arachidonic acid. This arachidonoyl transacylation system may play a very important role in the metabolism of lyso platelet-activating factor and also in the elimination or release of arachidonic acid from diacyl-GPC.  相似文献   

20.
Supplementation of growing MDCK canine kidney tubular epithelial cultures with linoleic acid produced a 3.6- to 4.9-fold increase in bradykinin-stimulated PGE2 release as measured by radioimmunoassay. Under these conditions the cell phospholipids contained 3.9-times more linoleic acid and 5.6-times more arachidonic acid, with the inositol, ethanolamine and choline phosphoglycerie fractions becoming enriched in arachidonic acid. By contrast, supplementation with arachidonic acid did not enhance bradykinin-stimulated PGE2 release even though the arachidonic acid content of the cell phospholipids was increased 8.8-fold. The distribution of radioactive prostaglandin products was unchanged by these fatty acid enrichments, with PGE2 accounting for 55 to 68% of the total output from [1-14C]arachidonic acid. Linoleic acid supplementation also produced a 2.5-fold increase in PGE2 formation stimulated by extracellular arachidonic acid, whereas supplementation during culture with arachidonic acid caused a 55 to 80% inhibition. This difference cannot be accounted for by changes in the ability of the cells to incorporate extracellular arachidonic acid. it is suggested that at least some of the effects of linoleate supplementation on prostaglandin production are due to the resulting enrichment of the intracellular phospholipid substrate pools with arachidonic acid. In addition, it appears that prolonged exposure to arachidonic acid during culture has an overriding inhibitory effect on prostaglandin production even though the total cell lipids bocome highly enriched in arachidonate.  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