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 0 毫秒
1.
The release of arachidonic acid from thrombin-stimulated platelets can be attributed to the action of phospholipase A2 on membrane phospholipid. Previously, analysis of individual subclasses of phospholipid demonstrated that 1-acyl-2-[3H]arachidonoyl-sn-glycerophosphocholine and to a lesser degree 1-acyl-2-[3H]arachidonoyl-sn-glycerophosphoethanolamine were the main source of [3H]arachidonic acid in thrombin-stimulated cells. In the present work, 1,2-diacyl phospholipid subclasses were analyzed as 1,2-diacylglycerobenzoates by high-pressure liquid chromatography in order to analyze arachidonate release as mass changes in individual molecular species of phospholipid. Following thrombin stimulation (5 U/ml, 5 min, 37 degrees C) all arachidonoyl-containing molecular species of 1,2-diacyl-sn-glycerophosphocholine decreased in mass and [3H]arachidonate content by almost 50%, while those of 1,2-diacyl-sn-glycerophosphoethanolamine decreased by 20%. The mass change was substantial and indicated that these phospholipids are a major source of arachidonate in stimulated cells. No variation was seen in the other non-arachidonate-containing molecular species of either subclass. Thus, deacylation of membrane 1,2-diacylglycerophosphocholine and 1,2-diacylglycerophosphoethanolamine by phospholipase A2 is selective for those molecular species of phospholipid containing arachidonic acid, suggesting that a certain proportion of arachidonoyl-containing molecular species of phospholipid are compartmentalized with the platelet membrane proximal to the site of action of this enzyme. These studies demonstrate that the human platelet is a cell poised and specialized to release rapidly substantial amounts of arachidonic acid upon stimulation.  相似文献   

2.
alpha-Synuclein has been implicated in the pathogenesis of many neurodegenerative diseases, including Parkinson's disease and Alzheimer's disease. Although the function of alpha-synuclein remains largely unknown, recent studies have demonstrated that this protein can interact with phospholipids. To address the role of alpha-synuclein in neurodegenerative disease, we have investigated whether it binds phospholipase D (PLD) and affects PLD activity in human embryonic kidney (HEK)-293 cells overexpressing wild type alpha-synuclein or the mutant forms of alpha-synuclein (A53T, A30P) associated with Parkinson's disease. Tyrosine phosphorylation of alpha-synuclein appears to play a modulatory role in the inhibition of PLD, because mutation of Tyr(125) to Phe slightly increases inhibitory effect of alpha-synuclein on PLD activity. Treatment with pervanadate or phorbol myristate acetate inhibits PLD more in HEK 293 cells overexpressing alpha-synuclein than in control cells. Binding of alpha-synuclein to PLD requires phox and pleckstrin homology domain of PLD and the amphipathic repeat region and non-Abeta component of alpha-synuclein. Although biologically important, co-transfection studies indicate that the interaction of alpha-synuclein with PLD does not influence the tendency of alpha-synuclein to form pathological inclusions. These results suggest that the association of alpha-synuclein with PLD, and modulation of PLD activity, is biologically important, but PLD does not appear to play an essential role in the pathophysiology of alpha-synuclein.  相似文献   

3.
The relationship between calcium mobilization and phospholipase D (PLD) activation in response to E-series prostaglandins (PGEs) was investigated in human erythroleukemia cells. Intracellular free Ca2+ concentration ([Ca2+]i) was increased by PGE1 and PGE2 over the same concentration range at which PLD activation was seen. Pretreatment of cells with pertussis toxin greatly inhibited the PGE-stimulated increase in [Ca2+]i, implying that a G protein participates in the PGE receptor signaling process. The peak level and also the plateau level of Ca2+ mobilization stimulated by these prostaglandins were markedly decreased in Ca(2+)-depleted medium, indicating that both extracellular and intracellular Ca2+ stores contribute to the changes in [Ca2+]i. Likewise, activation of PLD by PGE1 and PGE2 was abolished by pertussis toxin pretreatment or incubation in Ca(2+)-depleted medium. U73122, a putative phospholipase C inhibitor, blocked both Ca2+ mobilization and PLD activation in PGE-stimulated cells. Furthermore, the intracellular loading of BAPTA, a Ca2+ chelator, inhibited both Ca2+ mobilization and PLD activation by PGE1 and PGE2 in a similar dose-dependent manner. Simultaneous measurement of [Ca2+]i and PLD activity in the same cell samples indicated that PLD activity increases as a function of [Ca2+]i in a similar fashion in cells stimulated either by PGEs or by the calcium ionophore ionomycin. Taken together, these findings suggest that a rise in [Ca2+]i is necessary for PGE-stimulated PLD activity in human erythroleukemia cells.  相似文献   

4.
Phospholipase D (PLD) activity is commonly elevated in response to mitogenic signals. We reported previously that although the transformed phenotype induced by v-Src was dependent upon Raf-1, the PLD activity induced by v-Src was independent of Raf-1. This observation suggested to us that Raf would not likely be an activator of PLD. However, upon examination of PLD activity in v-Raf-transformed cells, surprisingly, we found that PLD activity is elevated to levels that were even higher than that observed in v-Src-transformed cells. To characterize the mechanism of v-Raf-induced PLD activity, we examined the dependence of v-Raf-induced PLD activity upon protein kinase C (PKC) the small GTPases Ral and Rho, which have all been implicated in the activation of PLD. The v-Raf-induced PLD activity was inhibited by dominant negative mutants for both Ral and Rho. The dependence upon Ral was particularly surprising since Ral is a downstream target of Ras, which is an upstream activator of Raf. Depleting cells of PKC by long term phorbol ester treatment actually increased PLD activity in v-Raf-transformed cells, indicating that v-Raf-induced PLD activity is not dependent on PKC. These data describe a novel mechanism for PLD activation by v-Raf that is independent of PKC, but dependent upon both Ral and Rho GTPases.  相似文献   

5.
The activation of phospholipase D in murine B cell lymphoma A20 cells treated with anti-Fas monoclonal antibody has been investigated. Fas cross-linking resulted in a both dose- and time-dependent increases in phospholipase D activity. There was a nearly maximum saturated rise in phospholipase D activity at the dose of 200 ng/ml anti-Fas monoclonal antibody showing a fourfold increase within 3 h. Fas activation also caused an approximately twofold increase of phosphatidylcholine-specific phospholipase C activity and 1,2-diacylglycerol release, which could be blocked by 30 min pretreatment with the phosphatidylcholine-specific phospholipase C inhibitor D609 (50 microgram/ml). Pretreatment of D609 also effectively inhibited the translocation of protein kinase C betaI and betaII from the cytosol to the membrane and the activation of phospholipase D induced by Fas cross-linking, suggesting that 1, 2-diacylglycerol released from the cellular phosphatidylcholine pool through phosphatidylcholine-specific phospholipase C plays a major role in protein kinase C/phospholipase D activation. Anti-Fas monoclonal antibody failed to elicit phosphoinositide-specific phospholipase C activation and any changes in the intracellular Ca2+ level in A20 cells, indicating that the phosphoinositide-mediated pathway is not involved in this Fas signaling. Therefore, these results suggest that Fas-mediated phospholipase D activation may be a consequence of primary stimulation of phosphatidylcholine-specific phospholipase C and that phospholipase D may play a role in Fas cross-linking signaling downstream from phosphatidylcholine-specific phospholipase C.  相似文献   

6.
We have shown earlier that oxidant-induced activation of phospholipase D (PLD) in vascular endothelial cells (ECs) is regulated by protein tyrosine kinases. To further understand the regulation of oxidant-induced PLD activation, we investigated the role of Src kinase. Treatment of bovine pulmonary artery ECs (BPAECs) with a model oxidant, diperoxovanadate (DPV), at 5 microM concentration, for 30 min, stimulated PLD activity (four- to eightfold), which was attenuated by tyrosine kinase inhibitors and by Src kinase-specific inhibitors PP-1 and PP-2, in a dose- and time-dependent fashion. Furthermore, BPAECs exposed to DPV (5 microM) for 2 min showed activation of Src kinase as observed by increased tyrosine phosphorylation and autophosphorylation in Src immunoprecipitates, which was attenuated by PP-2. Src immunoprecipitates of cell lysates from control BPAECs exhibited PLD activity in cell-free preparations, which was Arf- and Rho-sensitive and was enhanced at 2 min of DPV (5 microM) treatment. Also, Western blots of Src immunoprecipitates of control cells revealed the presence of PLD(1) and PLD(2), suggesting the association of PLD with Src kinase under basal conditions. However, exposure of cells to DPV (5 microM) for 2 min enhanced the association of PLD(2) but not PLD(1) with Src. Western blotting of immunoprecipitates of PLD(1) and PLD(2) isoforms of control BPAECs revealed the presence of Src under basal conditions and exposure of cells to DPV (5 microM) for 2 min enhanced the association of PLD(2) with Src in PLD(2) immunoprecipitates. Transient expression of a dominant negative mutant of Src in BPAECs attenuated DPV- but not TPA-induced PLD activation. In cell-free preparations, Src did not phosphorylate either PLD(1) or PLD(2) compared to protein kinase Calpha or p38 mitogen-activated protein kinase. These data show for the first time a direct association of Src with PLD in ECs and regulation of PLD in intact cells.  相似文献   

7.
Recently it was reported that tumor-promoting phorbol esters stimulate the production of phosphatidylethanol (PEt) in lymphocytes through the activation of phospholipase D (PLD). However, it remains unclear whether this activation is mediated through protein kinase (PKC). The study reported here shows that tumor promoters 12-0-tetradecanoylphorbol-13-acetate (TPA), phorbol dibutyrate (PDBU), 12-deoxyphorbol-13-phenylacetate (DOPP), 12-deoxyphorbol-13-phenylacetate-20-acetate (DOPPA) and mezerin activated PLD, as measured by the formation of PEt, whereas Concanavalin A (ConA) had no effect. Inhibitors of PKC, sphingosine (2 x 10(-6) M - 5 x 10(-6) M), H-7, HA1004 (5 x 10(-7) - 5 x 10(-6) M) and K252a (1 x 10(-7) - 1 x 10(-6) M) failed to block the PEt synthesis induced by TPA. In fact, sphingosine increased it. Other PKC activators, 1-oleoyl-2-acetylglycerol (OAG) and dioctanoylglycerol (DiC8) had no effect on lymphocyte PLD activity. Analysis of the phospholipid contents after stimulation by TPA showed that only phosphatidylcholine (PC) was significantly decreased. Interestingly, TPA activated PLD in intact cells but not in lysates or subcellular fractions. These observations suggest that stimulation of PLD-catalyzed PEt synthesis by TPA is not solely mediated through PKC activation.  相似文献   

8.
The activation of phospholipase D by platelet-activating factor (PAF) in the human promonocytic cell line U937 has been investigated. In cells prelabeled with [3H]palmitic acid, addition of PAF or phorbol 12-myristate 13-acetate (PMA) induced the synthesis of [3H]phosphatidylethanol, indicating phospholipase D activation. When U937 cells were preincubated for 5 min with PMA, and then stimulated with PAF, formation of phosphatidylethanol was greatly enhanced. In contrast, under the same experimental conditions PMA treatment blocked completely the PAF-induced inositol phosphates formation in cells prelabeled with [3H]inositol. Thus, PMA treatment demonstrates that phospholipase D activation can occur independently from phosphoinositide-specific phospholipase C activation during PAF stimulation in U937 cells. On the other hand, the data herein presented suggest that influx of external calcium is required for phospholipase D activation by PAF, as assessed by complete inhibition of the enzyme activity by chelation of extracellular calcium or by treatment with the calcium channel blocker verapamil. Based on these findings, a hypothetical model for phospholipase D activation is discussed.  相似文献   

9.
In order to test if phospholipase D (PLD) activity exists in the rat parotid gland, we took advantage of the fact that, in the presence of ethanol, PLD generates phosphatidylethanol (PEth) via a transphosphatidylation reaction. Lipid extracts of parotid acini prelabelled with [3H]myristic acid were analyzed by thin layer chromatography to determine [3H]phosphatidylethanol ([3H]PEth) formation. Carbamylcholine (1 mM) stimulated [3H]PEth formation in the presence of 2% ethanol, this effect was completely inhibited by atropine (10 microM). PMA (0.1-1 microM) and ionomycine (10 microM) also caused [3H]PEth generation. We conclude that a phospholipase D activity is present in the rat parotid gland and is regulated by muscarinic cholinergic receptors. Protein kinase C and calcium could also modulate this activity. This report provides the first evidence for the existence and receptor-linked regulation of phospholipase D in an exocrine gland, the rat parotid gland.  相似文献   

10.
We have previously reported that Fas cross-linking resulted in an increase in phospholipase D activity in A20 murine cells (J.-S. Han et al., Arch. Biochem. Biophys. 367, 233-239, 1999). In an attempt to explore the Fas downstream factor contributing to the activation of phospholipase D, we have investigated the possible involvement of a small GTP biding protein Ras in signaling events that were triggered by Fas cross-linking. Upon adenoviral expression of dominant negative mutant of Ras (N17Ras), an increase in phospholipase D activity by anti-Fas monoclonal antibody was diminished. Also, the Fas downstream signaling events triggered by Fas cross-linking such as the activation of phosphatidylcholine-specific phospholipase C, the increase in diacylglycerol level, and the translocation of protein kinase C to membrane fraction were all reduced by N17Ras expression. When parallel experiments were performed with manumycin-A, a Ras farnensyltransferase inhibitor, almost identical inhibitory effects on Fas downstream signaling were exhibited. These data suggest that Ras GTPase is essential in transmitting phospholipase D activation signal induced by Fas cross-linking and is located at phosphatidylcholine-specific phospholipase C upstream in Fas signaling cascades.  相似文献   

11.

Background  

Numerous investigations demonstrate a novel role of thyroid hormone as a modulator of signal transduction. Protein kinase C (PKC) is critical to the mechanism by which thyroid hormones potentiate both the antiviral and immunomodulatory actions of IFNγ in different cells and regulate the exchange of signalling phospholipids in hepatocytes. Because nothing is known about accumulation of PKC modulator - diacylglycerol in cells treated with T4, we examined the nongenomic effect of thyroid hormones on DAG formation and phospholipase activation in liver cells.  相似文献   

12.

Background

Many eukaryotes, including plants and fungi make spores that resist severe environmental stress. The micro-organism Dictyostelium contains a single phospholipase C gene (PLC); deletion of the gene has no effect on growth, cell movement and differentiation. In this report we show that PLC is essential to sense the environment of food-activated spores.

Results

Plc-null spores germinate at alkaline pH, reduced temperature or increased osmolarity, conditions at which the emerging amoebae can not grow. In contrast, food-activated wild-type spores return to dormancy till conditions in the environment allow growth. The analysis of inositol 1,4,5-trisphosphate (IP3) levels and the effect of added IP3 uncover an unexpected mechanism how PLC regulates spore germination: i) deletion of PLC induces the enhanced activity of an IP5 phosphatase leading to high IP3 levels in plc-null cells; ii) in wild-type spores unfavourable conditions inhibit PLC leading to a reduction of IP3 levels; addition of exogenous IP3 to wild-type spores induces germination at unfavourable conditions; iii) in plc-null spores IP3 levels remain high, also at unfavourable environmental conditions.

Conclusions

The results imply that environmental conditions regulate PLC activity and that IP3 induces spore germination; the uncontrolled germination of plc-null spores is not due to a lack of PLC activity but to the constitutive activation of an alternative IP3-forming pathway.  相似文献   

13.
Kim JH  Kim SW  Jung PJ  Yon C  Kim SC  Han JS 《Molecules and cells》2002,13(2):272-280
Liquid nitrogen (LN2) infusions are currently used in a slow controlled-rate freezing during cryopreservation. The effects of two different LN2 infusion frequencies (conventional, slow 50 infusions/min and high 120 infusions/min) were studied with frozen-thawed two-cell mouse embryos and their subsequent development to blastocysts. The embryos that were subjected to the high frequency LN2 infusion (HFLI) showed a significantly higher survival rate over the low frequency LN2 infusion (LFLI) (50.7 vs. 34.6%, P < 0.05). The blastocyst formation was also higher in HFLI (76.7%) than LFLI (44.0%, P < 0.05) with respective to the number of cells in a blastocyst of 71.6 8.0 (n = 20) and 62.5 +/- 4.7 (n = 20) (P < 0.05). The relative amount of H2O2 in an embryo that was assessed by a fluorescence intensity of 2',7'-dichlorofluorecein (DCF) showed a difference between the procedures with 16.6 +/- 1.6 (n = 21) and 23.4 +/- 1.8 (n = 24) for HFLI and LFLI, respectively (P < 0.05). Mitochondrial staining by Rhodamine 123 showed that the number and distribution of viable mitochondria were similar in both procedures, but fewer mitochondria were observed with a marked aggregation in the arrested embryos, indicating a mitochondrial disintegration. The mitochondrial membrane potential was visualized by a membrane potential-sensitive fluorescent probe, 5,5',6,6'-tetrachloro-1,1',3,3'-tetraethylbenzimidazolcarbocyanine iodide (JC-1). There was a decrease in the number of mitochondria that had a high membrane potential, and they showed a peripheral redistribution along the cell membrane in LFLI. A fluorescent staining of the actin filaments revealed a discontinuity that was noticeably at the peripheral "actin band" in LFLI. The DNA fragmentation was assessed by the dUTP nick end-labeling (TUNEL). The results showed a higher DNA fragmentation of blastocyst nuclei in LFLI compared to HFLI (65.6 vs. 36.0%, P < 0.05). Based on these observations, it was concluded that HFLI was better than LFLI in the case of freezing the mouse 2-cell embryos for preserving cytoskeletons and mitochondrial integrities. This could subsequently lead to a higher survival and developmental rate of the cryopreserved mouse embryos.  相似文献   

14.
12-O-Tetradecanoylphorbol-13-acetate (TPA) stimulated the release of [3H]ethanolamine from HeLa cells prelabeled with [3H]ethanolamine within 2 min, and of [3H]choline from cells prelabeled with [3H]choline after a lag of 10-20 min. This result suggests that TPA activates phospholipase D. Propranolol alone or propranolol plus TPA stimulated phosphatidic acid (PA) labeling in cells prelabeled with [3H]hexadecanol. In the presence of ethanol, TPA stimulated the accumulation of labeled phosphatidylethanol (PEth); no PEth was formed in the absence of TPA. TPA-dependent PEth accumulation was not observed in cells pretreated with TPA to down-regulate protein kinase C, whereas propranolol-induced accumulation of PA was unaffected by TPA pretreatment. Incubation of prelabeled cells with propranolol alone caused a rapid loss of label and phospholipid mass from both phosphatidylethanolamine and phosphatidylcholine (PC) together with an accumulation of PA and phosphatidylinositol plus phosphatidylserine. When [3H]hexadecanol-prelabeled cells were pulse labeled with 32P to label nucleotide pools, propranolol induced the accumulation of both 3H- and 32P-labeled PA. When cells were prelabeled with lyso-PC double labeled with 3H and 32P, and incubated with propranolol, only 3H-labeled PA accumulated, indicating that the pathways involved in the basal turnover of PC resulted in the loss of 32P from the lipid. These results suggest that the basal turnover of phosphatidylethanolamine and PC involves the sequential actions of phospholipase C, diglyceride kinase, and PA phosphohydrolase.  相似文献   

15.
The IFN-gamma linked PKC activation in endothelial cells was analysed. It was shown that IFN-gamma activates PKC in three transient and separate cycles within the first 60 minutes after IFN-gamma stimulation. Before each PKC activation there was an increase in DAG level. IP3, phosphocholine and choline productions were measured to determine the origin of DAG. Neither of the PLC products, IP3 or phosphocholine, were released after IFN-gamma stimulation. On the other hand the PLD products choline and PA were released before all the three activation cycles of PKC.  相似文献   

16.
To determine the role of protein tyrosine phosphorylation in the activation of phospholipase D (PLD), electropermeabilized HL-60 cells labeled in [3H]alkyl-phosphatidylcholine were treated with vanadate derivatives. Micromolar concentrations of vanadyl hydroperoxide (V(4+)-OOH) induced accumulation of tyrosine-phosphorylated proteins. Concomitantly, V(4+)-OOH or a combination of vanadate and NADPH elicited a concentration- and time-dependent accumulation of phosphatidic acid (PtdOH). In the presence of ethanol a sustained formation of phosphatidylethanol was observed, indicating that a type D phospholipase was activated. A good correlation was found to exist between the accumulation of tyrosine-phosphorylated proteins and activation of PLD. The V(4+)-OOH concentration dependence of the two responses was nearly identical, and the time course of activation was similar, with tyrosine phosphorylation preceding PLD activation by approximately 1 min. The ability of V(4+)-OOH to induce both responses was found to be strictly dependent on the presence of ATP and/or Mg2+, suggesting that PLD activation involves phosphotransferase reactions. Accordingly, ST638, a tyrosine kinase inhibitor, reduced concomitantly tyrosine phosphorylation and PLD activation elicited by V(4+)-OOH. The mechanism of action of V(4+)-OOH was investigated. The diacylglycerol kinase inhibitors, dioctanoylethylene glycol and R59022 potentiated PLD stimulation by exogenous diacylglycerol but not by V(4+)-OOH. Moreover, stimulation by V(4+)-OOH and by phorbol esters was synergystic. Therefore, diacylglycerol-induced activation of protein kinase C is unlikely to mediate the effects of V(4+)-OOH. The response of PLD to V(4+)-OOH was larger than that to guanosine 5'-(gamma-thio)triphosphate. Moreover, the effects of GTP gamma S and V(4+)-OOH were additive. Hence, activation of G proteins cannot account for the stimulation of PLD by V(4+)-OOH. V(4+)-OOH also triggers a burst of O2 consumption by the NADPH oxidase. Inhibition of PtdOH accumulation by addition of ethanol or by ST638 abolished this respiratory burst. Together, the results establish a strong correlation between tyrosine phosphorylation, PLD activation, and stimulation of the NADPH oxidase in HL-60 cells, suggesting a causal relationship.  相似文献   

17.
18.
Activation of phospholipase D (PLD) and protein kinase C (PKC) as well as calcium mobilization are essential signals for degranulation of mast cells. However, the exact role of PLD in degranulation remains undefined. In this study we have tested the hypothesis that the PLD product, phosphatidic acid, and diacylglycerides generated therefrom might promote activation of PKC. Studies were conducted in two rodent mast cell lines that were stimulated with Ag via FcepsilonRI and a pharmacologic agent, thapsigargin. Diversion of production of phosphatidic acid to phosphatidylbutanol (the transphosphatidylation reaction) by addition of l-butanol suppressed both the translocation of diacylglyceride-dependent isoforms of PKC to the membrane and degranulation. Tertiary-butanol, which is not a substrate for the transphosphatidylation, had a minimal effect on PKC translocation and degranulation, and 1-butanol itself had no effect on PKC translocation when PKC was stimulated directly with phorbol ester, 12-O-tetradecanoylphorbol-13-acetate. Also, in cells transfected with small inhibitory RNAs directed against PLD1 and PLD2, activation of PLD, generation of diacylglycerides, translocation of PKC, and degranulation were all suppressed. Phorbol ester, which did not stimulate degranulation by itself, restored degranulation when used in combination with thapsigargin whether PLD function was disrupted with 1-butanol or the small inhibitory RNAs. However, degranulation was not restored when cells were costimulated with Ag and phorbol ester. These results suggested that the production of phosphatidic acid by PLD facilitates activation of PKC and, in turn, degranulation, although additional PLD-dependent processes appear to be critical for Ag-mediated degranulation.  相似文献   

19.
20.
TPA, a potent PKC activator, inhibits myogenic differentiation and activates phospholipase D (PLD). We evaluated the involvement of PLD in the TPA effects on L6 myoblasts differentiation. TPA, at concentrations inhibiting differentiation of L6 cells, induced a strong, though transient, PLD activation. Surprisingly, at nanomolar concentration, TPA induced both myogenic differentiation and sustained activation of PLD. Differential effect of TPA can be ascribed to PKC downregulation induced by highest TPA concentrations. TPA-induced differentiation was inhibited by 1-butanol, confirming the involvement of PLD in this effect. These data suggest that prolonged elevation of PLD activity is required for myogenic differentiation.  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