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 15 毫秒
1.
The present study was conducted to examine the role of a major cardiac phospholipase C (PLC) isozyme, PLC-gamma 1, in cardiomyocytes during oxidative stress. Left ventricular cardiomyocytes were isolated by collagenase digestion from adult male Sprague-Dawley rats (250-300 g) and treated with 20, 50, and 100 microM H2O2 for 15 min. A concentration-dependent (up to 50 microM) increase in the mRNA level and membrane protein content of PLC-gamma 1 was observed with H2O2 treatment. Furthermore, PLC-gamma 1 was activated in response to H2O2, as revealed by an increase in the phosphorylation of its tyrosine residues. There was a marked increase in the phosphorylation of the antiapoptotic protein Bcl-2 by H2O2; this change was attenuated by a PLC inhibitor, U-73122. Although both protein kinase C (PKC)-delta and -epsilon protein contents were increased in the cardiomyocyte membrane fraction in response to H2O2, PKC-epsilon activation, unlike PKC-delta, was attenuated by U-73122 (2 microM). Inhibition of PKC-epsilon with inhibitory peptide (0.1 microM) prevented Bcl-2 phosphorylation. Moreover, different concentrations (0.05, 0.1, and 0.2 microM) of this peptide augmented the decrease in cardiomyocyte viability in response to H2O2. In addition, a decrease in cardiomyocyte viability, as assessed by trypan blue exclusion, due to H2O2 was also seen when cells were pretreated with U-73122 and was as a result of increased apoptosis. It is therefore suggested that PLC-gamma 1 may play a role in cardiomyocyte survival during oxidative stress via PKC-epsilon and phosphorylation of Bcl-2.  相似文献   

2.
The secretagogue effect of prolactin (PRL) on casein release by epithelial mammary cells has been previously related to stimulation of the phospholipase A2-arachidonic acid cascade. In order to determine whether other intracellular pathways are implicated in this secretagogue effect, different agents acting on protein kinase C (PKC) and phospholipase C (PLC) activity have been assessed in vitro in lactating rabbit mammary gland fragments. Phorbol ester (20 nm TPA and 1-oleoyl-2-acetyl-sn-glycerol (10 microM (OAG) stimulated newly synthesized casein secretion and potentiated the PRL secretatogue effect. However, 100 microM quercetin, 100 microM H-7 and 5 and 20 nM staurosporine did not inhibit the latter effect. Exogenous PLC did not stimulate casein secretion. PRL did not affect production of inositol phosphates (IPs) during 10 or 60 min exposure. These results show that PKC activation may increase basal levels of casein secretion, and demonstrate that PRL does not act primarily via PKC activation or by PLC activation to stimulate casein secretion.  相似文献   

3.
We have examined signal transduction via membrane IgM (mIgM) in resting and cycling human B cells. Crosslinking mIgM on all of the cell types studied transduced a signal through the phosphatidylinositol pathway, producing inositol 1,4,5-trisphosphate and release of intracellular free calcium. These second messengers were formed regardless of quantitative or qualitative differences in the surface expression of mIgM: cells that had low levels of surface IgM (T-51) or had no light chain associated with surface heavy chain (DB) signaled phosphatidylinositol pathway activation after mIgM crosslinking. Production of specific lipid products in nonquiescent B cells differed from that in normal resting cells. Ligation of surface immunoglobulin on resting B cells resulted in sustained increases of both diacylglycerol and phosphatidic acid, two lipids that can influence PKC activation. Whereas PKC was strongly activated in normal tonsillar B cells, several cell lines had reduced PKC activation following crosslinking of mIgM. The reduction in protein kinase C activation correlated with the absence or reduced levels of phosphatidic acid or diacylglycerol following stimulation: protein kinase C translocated and was activated only in cells that had elevated levels of both diacylglycerides and phosphatidic acid. Anti-IgM-induced phosphorylation of a protein kinase C substrate protein CD20, also increased in those cells having PKC activation and not in cells in which kinase activity was reduced. CD20 phosphorylation also increased following the direct addition of exogenous phosphatidic acid to resting B cells. Together, these observations show that the generation of lipid products following mIgM crosslinking in resting cells can vary from that in cycling cells and may relate to the different levels of PKC activation. In a companion study we report that ligation of surface IgM activates both an acyltransferase and phospholipase D to form phosphatidic acid.  相似文献   

4.
Rod outer segments (ROS) exhibit high acyltransferase (AT) activity, the preferred substrate of which being lysophosphatidylcholine. To study factors possibly regulating ROS AT activity purified ROS membranes were assayed under conditions under which protein kinase C (PKC), cAMP-dependent protein kinase (PKA), and phosphatases were stimulated or inhibited. PKC activation produced a significant increase in the acylation of phosphatidylethanolamine (PE) and phosphatidylinositol (PI) with oleate, it inhibited phosphatidylcholine (PC) acylation, and phosphatidylserine (PS) and phosphatidic acid (PA) acylation remained unchanged. ROS PKA activation resulted in increased oleate incorporation into PS and PI while the acylation of PC, PE, and PA remained unchanged. Inhibition of ROS PKC or PKA produced, as a general trait, inverse effects with respect to those observed under kinase-stimulatory conditions. ROS phosphatase 2A was inhibited by using okadaic acid, and the changes observed in AT activity are described. These findings suggest that changes in ROS protein phosphorylation produce specific changes in AT activity depending on the phospholipid substrate. The effect of light on AT activity in ROS membranes was also studied and it is reported that acylation in these membranes remains unchanged independent of the illumination condition used.  相似文献   

5.
In this study, the underlying mechanisms of stimulation by cyclocommunin, a natural pyranoflavonoid, of respiratory burst in rat neutrophils was investigated. Cyclocommunin evoked a concentration-dependent stimulation of superoxide anion (O2*-) generation with a slow onset and long lasting profile. The maximum response (16.4+/-2.3 nmol O2*-/10 min per 10(6) cells) was observed at 3-10 microM cyclocommunin. Cyclocommunin did not activate NADPH oxidase in a cell-free system. Cells pretreated with pertussis toxin or n-butanol did not affect the cyclocommunin-induced O2*- generation. However, a protein kinase inhibitor staurosporine and EGTA greatly reduced the O2*-generation caused by cyclocommunin. Treatment of neutrophils with phorbol 12-myristate 13-acetate (PMA), but not with formylmethionyl-leucyl-phenylalanine (fMLP), for 20 min significantly reduced the O2*- generation following the subsequent stimulation of cells with cyclocommunin. Cyclocommunin did not affect the cellular mass of phosphatidic acid (PA). Neither the tyrosine kinase inhibitor, genistein, nor the p38 mitogen-activated protein kinase (MAPK) inhibitor, SB203580, affected cyclocommunin-induced O2*- generation. The enzyme activities of neutrophil cytosolic and membrane-associated protein kinase C (PKC) were both increased significantly with 100 microM cyclocommunin. The membrane-associated PKC-theta and PKC-beta were increased following the stimulation of neutrophils with 30 and 100 microM cyclocommunin, respectively. Cyclocommunin reduced the [3H]phorbol 12,13-dibutyrate ([3H]PDB) binding to cytosolic PKC in a concentration-dependent manner. Cyclocommunin (> or =3 microM) significantly evoked a slow and long lasting [Ca2+]i elevation in neutrophils, and a phospholipase C (PLC) inhibitor U73122 greatly inhibited these Ca2+ responses. Moreover, the increase in cellular inositol bis- and trisphosphate (IP2 and IP3) levels were observed in neutrophils stimulated with 30 microM cyclocommunin for 3 min. Collectively, these results indicate that the stimulation of respiratory burst by cyclocommunin is probably mediated by the synergism of PKC activation and [Ca2+]i elevation in rat neutrophils.  相似文献   

6.
The possible involvement of protein kinase C activation in regulating PAF-stimulated PLC activity was studied in rabbit platelets. PAF (100 nM for 5 seconds) stimulated incorporation of 32P into proteins and caused [3H]InsP3 levels to increase about 260% of control. These responses were compared after platelets were pretreated with either PAF, phorbol 12-myristate 13-acetate (PMA) or staurosporine and also after pretreatments with staurosporine followed by PAF or PMA. Pretreating platelets with staurosporine potentiated PAF-stimulated [3H]InsP3 levels by 54% and blocked protein phosphorylation. Pretreatments with PAF and PMA caused PAF-stimulated [3H]InsP3 levels to decrease to 115 and 136%, respectively. Staurosporine pretreatment blocked the decrease caused by the PMA pretreatment but not that by PAF. This study demonstrates that PAF-stimulated PLC activity is negatively affected by protein kinase C (PKC) activation and that inhibition of PKC activity did not prevent desensitization of PLC by PAF.  相似文献   

7.
The diacylglycerol kinase inhibitor R59022 (10 microM) potentiates secretion and aggregation responses in human platelets challenged with sub-maximal concentrations of thrombin. Potentiation correlates closely with increased formation of diacylglycerol, increased phosphorylation of a 40 kDa protein, a known substrate for protein kinase C, and with decreased formation of phosphatidic acid, the product of diacylglycerol kinase. Phosphorylation of myosin light chains, formation of inositol phosphates and the mobilization of Ca2+ by thrombin are not affected by R59022 (10 microM). These data support a role for protein kinase C in platelet aggregation and secretion, and provide further evidence that endogenous diacylglycerols bring about the activation of this enzyme. These data also add further argument against a role for phosphatidic acid in platelet activation.  相似文献   

8.
Calf spleen profilin is shown to be an in vitro substrate of purified human placental protein kinase C (PKC), with an apparent Km of 4 microM. Phosphatidylinositol bisphosphate (PIP2) was an effective activator of the profilin phosphorylation by PKC and caused a maximum 13-fold increase of Vmax with a half maximal effect at 40 micrograms/ml. The action of PIP2 was not mimicked by phosphatidylserine, phosphatidic acid or phosphatidylinositol, whereas phosphatidylinositol monophosphate was slightly stimulatory. By contrast, protein kinase C-dependent phosphorylation of histone type III-S, myelin basic protein or lipocortin-I was not affected by PIP. It is suggested that PIP2 modifies the nature of the profilin-PKC interactions.  相似文献   

9.
A potential role of arachidonic acid in the modulation of insulin secretion was investigated by measuring its effects on calmodulin-dependent protein kinase and protein kinase C in islet subcellular fractions. The results were interpreted in the light of arachidonic acid effects on insulin secretion from intact islets. Arachidonic acid could replace phosphatidylserine in activation of cytosolic protein kinase C (K0.5 of 10 microM) and maximum activation was observed at 50 microM arachidonate. Arachidonic acid did not affect the Ca2+ requirement of the phosphatidylserine-stimulated activity. Arachidonic acid (200 microM) inhibited (greater than 90%) calmodulin-dependent protein kinase activity (K0.5 = 50-100 microM) but modestly increased basal phosphorylation activity (no added calcium or calmodulin). Arachidonic acid inhibited glucose-sensitive insulin secretion from islets (K0.5 = 24 microM) measured in static secretion assays. Maximum inhibition (approximately 70%) was achieved at 50-100 microM arachidonic acid. Basal insulin secretion (3 mM glucose) was modestly stimulated by 100 microM arachidonic acid but in a non-saturable manner. In perifusion secretion studies, arachidonic acid (20 microM) had no effect on the first phase of glucose-induced secretion but nearly completely suppressed second phase secretion. At basal glucose (4 mM), arachidonic acid induced a modest but reproducible biphasic insulin secretion response which mimicked glucose-sensitive secretion. However, phosphorylation of an 80 kD protein substrate of protein kinase C was not increased when intact islets were incubated with arachidonic acid, suggesting that the small increases in insulin secretion seen with arachidonic acid were not mediated by protein kinase C. These data suggest that arachidonic acid generated by exposure of islets to glucose may influence insulin secretion by inhibiting the activity of calmodulin-dependent protein kinase but probably has little effect on protein kinase C activity.  相似文献   

10.
Familial HDL deficiency (FHD) is the heterozygous form of Tangier disease (TD). Mutations of the ABCA1 gene cause FHD and TD. FHD/TD cells are unable to normally efflux cholesterol onto nascent HDL particles, which are rapidly catabolized. TD fibroblasts have an abnormal pattern of PLC and PLD activation following cell stimulation with HDL(3) or apolipoprotein A-I (apoA-I). We examined cellular cholesterol efflux in FHD and TD fibroblasts by phospholipid-derived-molecules, compared with that of normal cells. We used the PKC agonist 1,2-dioctanoylglycerol (DOG) and phorbol myristate acetate (PMA) to activate PKC, calphostin C, and GO 6976, as inhibitors of PKC; phosphatidic acid (PA), which is the product of PLD, and lysophosphatidic acid (LPA), phosphatidylcholine, sphingomyelin, and beta-cyclodextrin to investigate their potential effect in modulating cellular cholesterol efflux in [(3)H]cholesterol-labeled and cholesterol-loaded fibroblasts. Phosphatidylcholine, sphingomyelin, and beta-cyclodextrin promoted cholesterol efflux in an identical fashion in control, FHD, or TD fibroblasts. In a dose-dependent fashion, DOG (0-200 microM) increased apoA-I-mediated cellular cholesterol efflux by 40% in controls, 71% in FHD, and 242% in TD cells. PMA similarly increased cholesterol efflux to a maximum of 256% in controls, 182% in FHD, and 191% in TD cells. This effect was inhibited by calphostin C. PA (100 microM) also increased cholesterol efflux by 25% in control, 44% in FHD, and 100% in TD cells. Conversely, LPA reduced cholesterol efflux in a dose-dependent fashion in control and FHD cells (-50%, 200 microM) but not in TD cells, where efflux was increased by 140%. Propranolol (100 microM) significantly increased cholesterol efflux at 24 h in all three cell lines. n-Butanol partially decreased the DOG-mediated increase in cholesterol efflux. The inhibitory effect of calphostin C on DOG-stimulated cholesterol efflux could be partially overcome by propranolol, suggesting that PA is a downstream mediator of PKC-stimulated cholesterol efflux.We conclude that PLC and PLD activities are required for apoA-I-mediated cellular cholesterol efflux, and modulating cellular PA concentration can correct, at least partially, the cholesterol efflux defect in FHD and TD.  相似文献   

11.
The aim of this study was to investigate the involvement of calmodulin in phospholipase D activation in SH-SY5Y cells. Cells prelabelled with [3H]-palmitic acid were incubated with calmodulin antagonists and/or other compounds. Phosphatidylethanol, a specific marker for phospholipase D activity, and phosphatidic acid were analysed. The calmodulin antagonists, calmidazolium and trifluoperazine, induced an extensive increase in phosphatidylethanol formation, and thus increased basal phospholipase D activity, in a dose- and time-dependent manner. The effect of calmidazolium on carbachol-induced activation of muscarinic receptors was also studied. Calmidazolium did not significantly affect the amount of phosphatidylethanol formed following carbachol addition. However, taking into account the increase in basal activity observed after calmidazolium addition, calmidazolium probably inhibits the muscarinic receptor-induced phospholipase D activation. In addition to phosphatidylethanol, basal phosphatidic acid levels were also increased after calmidazolium and trifluoperazine addition. Incubation with calmidazolium (10 microM) for 10 min induced a two-fold increase in phosphatidic acid. The calmidazolium-induced increase in basal phospholipase D activity was not affected by the protein kinase inhibitors H7 and staurosporine. On the other hand tyrosine kinase inhibitors abolished the calmidazolium-induced activation of phospholipase D. Calmidazolium also induced tyrosine phosphorylation in parallel to the phospholipase D activation. In conclusion, our data indicate that calmodulin antagonists induce phospholipase D activity in SH-SY5Y cells via a tyrosine kinase dependent pathway. This may point to a negative control of phospholipase D by calmodulin although a calmodulin-independent mechanism cannot be excluded. Calmodulin antagonists may be useful tools to further elucidate the mechanisms of phospholipase D regulation.  相似文献   

12.
In inflammatory cells, agonist-stimulated arachidonic acid (AA) release is thought to be induced by activation of group IV Ca(2+)-dependent cytosolic phospholipase A(2) (cPLA(2)) through mitogen-activated protein kinase (MAP kinase)- and/or protein kinase C (PKC)-mediated phosphorylation and Ca(2+)-dependent translocation of the enzyme to the membrane. Here we investigated the role of phospholipases in N-formylmethionyl-l-leucyl-l-phenylalanine (fMLP; 1 nM-10 microM)-induced AA release from neutrophil-like db-cAMP-differentiated HL-60 cells. U 73122 (1 microM), an inhibitor of phosphatidyl-inositol-4,5-biphosphate-specific phospholipase C, or the membrane-permeant Ca(2+)-chelator 1, 2-bis?2-aminophenoxy?thane-N,N,N',N'-tetraacetic acid (10 microM) abolished fMLP-mediated Ca(2+) signaling, but had no effect on fMLP-induced AA release. The protein kinase C-inhibitor Ro 318220 (5 microM) or the inhibitor of cPLA(2) arachidonyl trifluoromethyl ketone (AACOCF(3); 10-30 microM) did not inhibit fMLP-induced AA release. In contrast, AA release was stimulated by the Ca(2+) ionophore A23187 (10 microM) plus the PKC activator phorbol myristate acetate (PMA) (0.2 microM). This effect was inhibited by either Ro 318220 or AACOCF(3). Accordingly, a translocation of cPLA(2) from the cytosol to the membrane fraction was observed with A23187 + PMA, but not with fMLP. fMLP-mediated AA release therefore appeared to be independent of Ca(2+) signaling and PKC and MAP kinase activation. However, fMLP-mediated AA release was reduced by approximately 45% by Clostridium difficile toxin B (10 ng/ml) or by 1-butanol; both block phospholipase D (PLD) activity. The inhibitor of phosphatidylcholine-specific phospholipase C (PC-PLC), D609 (100 microM), decreased fMLP-mediated AA release by approximately 35%. The effect of D609 + 1-butanol on fMLP-induced AA release was additive and of a magnitude similar to that of propranolol (0.2 mM), an inhibitor of phosphatidic acid phosphohydrolase. This suggests that the bulk of AA generated by fMLP stimulation of db-cAMP-differentiated HL-60 cells is independent of the cPLA(2) pathway, but may originate from activation of PC-PLC and PLD.  相似文献   

13.
Taxanes can induce drug resistance by increasing signaling pathways such as PI3K/Akt and ERK, which promote survival and cell growth in human cancer cells. We have previously shown that long chain n-3 polyunsaturated fatty acids, such as docosahexaenoic acid (DHA, 22:6n-3) decrease resistance of experimental mammary tumors to anticancer drugs. Our objective was to determine whether DHA could increase tumor sensitivity to docetaxel by down-regulating these survival pathways. In docetaxel-treated MDA-MB-231 cells, phosphorylated-ERK1/2 levels were increased by 60% in membrane and nuclear compartments, compared to untreated cells. Our data showed that ERK1/2 activation depended on PKC activation since: i) enzastaurin (a pan-PKC inhibitor) blocked docetaxel-induced ERK1/2 phosphorylation ii) docetaxel increased PKC activity by 30% and phosphatidic acid level by 1.6-fold iii) inhibition of PKCε and PKCδ by siRNA resulted in reduced phosphorylated ERK1/2 levels. In DHA-supplemented cells, docetaxel was unable to increase PKCε and δ levels in membrane and nuclear fractions, resulting in diminished ERK1/2 phosphorylation and increased docetaxel efficacy. Reduced membrane level of PKCε and PKCδ was associated with significant incorporation of DHA in all phospholipids, including phosphatidylcholine which is a major source of phosphatidic acid. Additionally, examination of the Akt pathway showed that DHA could repress docetaxel-induced Ser473Akt phosphorylation. In rat mammary tumors, dietary DHA supplementation during docetaxel chemotherapy repressed ERK and Akt survival pathways and in turn strongly improved taxane efficacy. The P-ERK level was negatively correlated with tumor regression. These findings are of potential clinical importance in treating chemotherapy-refractory cancer.  相似文献   

14.
15.
Feedback regulation of phospholipase C-beta by protein kinase C   总被引:9,自引:0,他引:9  
Treatment of a variety of cells and tissues with 12-O-tetradecanoylphorbol-13-acetate (TPA), an activator of protein kinase C (PKC) results in the inhibition of receptor-coupled inositol phospholipid-specific phospholipase C (PLC) activity. To determine whether or not the targets of TPA-activated PKC include one or more isozymes of PLC, studies were carried out with PC12, C6Bu1, and NIH 3T3 cells, which contain at least three PLC isozymes, PLC-beta, PLC-gamma, and PLC-delta. Treatment of the cells with TPA stimulated the phosphorylation of serine residues in PLC-beta, but the phosphorylation state of PLC-gamma and PLC-delta was not changed significantly. Phosphorylation of bovine brain PLC-beta by PKC in vitro resulted in a stoichiometric incorporation of phosphate at serine 887, without any concomitant effect on PLC-beta activity. We propose, therefore, that rather than having a direct effect on enzyme activity, the phosphorylation of PLC-beta by PKC may alter its interaction with a putative guanine nucleotide-binding regulatory protein and thereby prevent its activation.  相似文献   

16.
The activation of protein kinase C was investigated in digitonin-permeabilized human neuroblastoma SH-SY5Y cells by measuring the phosphorylation of the specific protein kinase C substrate myelin basic protein4-14. The phosphorylation was inhibited by the protein kinase C inhibitory peptide PKC19-36 and was associated to a translocation of the enzyme to the membrane fractions of the SH-SY5Y cells. 1,2-Dioctanoyl-sn-glycerol had no effect on protein kinase C activity unless the calcium concentration was raised to concentrations found in stimulated cells (above 100 nM). Calcium in the absence of other activators did not stimulate protein kinase C. Phorbol 12-myristate 13-acetate was not dependent on calcium for the activation or the translocation of protein kinase C. The induced activation was sustained for 10 min, and thereafter only a small net phosphorylation of the substrate could be detected. Calcium or dioctanoylglycerol, when applied alone, only caused a minor translocation, whereas in combination a marked translocation was observed. Arachidonic acid (10 microM) enhanced protein kinase C activity in the presence of submaximal concentrations of calcium and dioctanoylglycerol. Quinacrine and p-bromophenacyl bromide did not inhibit calcium- and dioctanoylglycerol-induced protein kinase C activity at concentrations which are considered to be sufficient for phospholipase A2 inhibition.  相似文献   

17.
Ethanol and diolein stimulate PKC translocation in astroglial cells   总被引:3,自引:0,他引:3  
S Skwish  W Shain 《Life sciences》1990,47(12):1037-1042
Ethanol exposure stimulates taurine release from astroglial cells. To determine if ethanol mediates this release using protein kinase C (PKC), PKC activity was measured using LRM55 astroglial cells. When ethanol (25-200 mM) or diolein (3 microM) was applied to cells for 30 seconds, PKC activity was observed to decrease in the cytosol and increase in the membrane fraction of the cell while the whole cell activity remained unchanged. The membrane-associated activity increased by almost 100%. When ethanol (100 mM) and diolein (3 microM) were applied simultaneously, membrane-associated activity increased to become 3-5 times greater than when either PKC activator was applied alone. These changes in PKC activity parallel changes in taurine release observed when cells are exposed to ethanol and the PKC activator diolein. Ethanol-stimulated release may be associated with the translocation of PKC activity from the cytosol to the membrane.  相似文献   

18.
The phenolic antioxidant 2,6-bis(1,1-dimethyl ethyl)-4-methylphenol (BHT) evokes a transient phosphorylation of two platelet proteins of Mr 20,000 and 47,000 that are well-known substrates of protein kinase C (PKC) and, similarly to phorbol esters, a slight but persistent phosphorylation of a protein of Mr 26,000. These effects are observed both in the presence and in the absence of extracellular calcium, but are abolished in the presence of the protein kinase C inhibitor staurosporine. The phosphorylation of the 47 kDa protein takes place mostly at the serine and, to a lesser extent, at threonine residues. BHT induces an increased binding of tritiated phorbol dibutyrate to platelets indicating a PKC translocation from cytosol to plasma membrane. Addition of BHT (20 microM) a few min prior to thrombin causes inhibition of both agonist-evoked protein phosphorylation and increase in the Ca2+ concentration, the latter inhibition being counteracted by staurosporine. The inhibitory effect lasts for several minutes even after removal of BHT from the cellular suspending medium. Similar results are obtained with nordihydroguaiaretic acid, whereas 2- and 3-tert-butyl-4-methoxyphenol (BHA) produce only slight effects. BHT activates the protein kinase C purified from pig brain in a concentration-dependent manner (up to 200 microM), whereas it does not affect the activity of other purified protein kinases such as type 1 and 2 casein kinases, type II A, II B and III tyrosine protein kinases from rat spleen and the catalytic subunit of cyclic AMP-dependent protein kinase. It is concluded that, similarly to diacylglycerols and phorbol esters, these phenolic antioxidants activate the protein kinase C, which in turn desensitizes platelets towards subsequent phospholipase C activation.  相似文献   

19.
We made stable cell lines overexpressing PLD1 (GP-PLD1) from GP+envAm12 cell, a derivative of NIH 3T3 cell. PLD1 activity and extracellular signal-regulated kinase (ERK) phosphorylation were enhanced in GP-PLD1 cells by the treatment of lysophosphatidic acid (LPA). In contrast, these LPA-induced effects were attenuated with the pretreatment of pertussis toxin (PTX) or protein kinase C (PKC) inhibitor. Moreover, accumulation of phosphatidic acid (PA), a product of PLD action, potentiated the LPA-induced ERK activation in GP-PLD1 cells while blocking of PA production with the treatment of 1-butanol attenuated LPA-induced ERK phosphorylation. From these results, we suggest that LPA activate PLD1 through pertussis toxin-sensitive G protein and PKC-dependent pathways, then PA produced from PLD1 activation facilitate ERK phosphorylation.  相似文献   

20.
Phosphatidic acid has been proposed to contribute to the mitogenic actions of various growth factors. In32P-labeled neonatal rat cardiac fibroblasts, 100 nM [Sar1]angiotensin II was shown to rapidly induce formation of32P-phosphatidic acid. Levels peaked at 5 min (1.5-fold above control), but were partially sustained over 2 h. Phospholipase D contributed in part to phosphatidic acid formation, as32P- or3H-phosphatidylethanol was produced when cells labeled with [32P]H3PO4 or 1-O-[1,2-3H]hexadecyl-2-lyso-sn-glycero-3-phosphocholine were stimulated in the presence of 1% ethanol. [Sar1]angiotensin II-induced phospholipase D activity was transient and mainly mediated through protein kinase C (PKC), since PKC downregulation reduced phosphatidylethanol formation by 68%. Residual activity may have been due to increased intracellular Ca2+, as ionomycin also activated phospholipase D in PKC-depleted cells. Phospholipase D did not fully account for [Sar1]angiotensin II-induced phosphatidic acid: 1) compared to PMA, a potent activator of phospholipase D, [Sar1]angiotensin II produced more phosphatidic acid relative to phosphatidylethanol, and 2) PKC downregulation did not affect [Sar1]angiotensin II-induced phosphatidic acid formation. The diacylglycerol kinase inhibitor R59949 depressed [Sar1]angiotensin II-induced phosphatidic acid formation by only 21%, indicating that activation of a phospholipase C and diacylglycerol kinase also can not account for the bulk of phosphatidic acid. Thus, additional pathways not involving phospholipases C and D, such asde novo synthesis, may contribute to [Sar1]angiotensin II-induced phosphatidic acid in these cells. Finally, as previously shown for [Sar1]angiotensin II, phosphatidic acid stimulated mitogen activated protein (MAP) kinase activity. These results suggest that phosphatidic acid may function as an intracellular second messenger of angiotensin II in cardiac fibroblasts and may contribute to the mitogenic action of this hormone on these cells. (Mol Cell Biochem141: 135–143, 1994)Abbreviations DAG diacylglycerol - DMSO dimethyl sulfoxide - lysoPC 1-O-hexadecyl-2-lyso-sn-glycero-3-phosphocholine - NRCF newborn rat cardiac fibroblasts - PA phosphatidic acid - PAPase phosphatidic acid phosphohydrolase - PC phosphatidylcholine - PEt phosphatidylethanol - PI phosphatidylinositol - PL (labeled) phospholipids - PLC phospholipase C - PLD phospholipase D Drs. G. W. Booz and M. M. Taher contributed equally to the work described here.  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