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 468 毫秒
1.
We have previously reported that Fas-resistant A20 cells (FasR) have phospholipase D (PLD) activity upregulated by endogenous PLD2 overexpression. In the present study, we investigated how overexpressed PLD2 in FasR could generate survival signals by regulating the protein levels of anti-apoptotic Bcl-2 and Bcl-xL. To confirm the effect of PLD2 on Bcl-2 protein levels, we transfected PLD2 into wild-type murine B lymphoma A20 cells. The transfected cells showed markedly the increases in Bcl-2 and Bcl-xL protein levels, and became resistant to Fas-induced apoptosis, similar to FasR. Treatment of wild-type A20 cells with phosphatidic acid (PA), the metabolic end product of PLD2 derived from phosphatidylcholin, markedly increased levels of anti-apoptotic Bcl-2 and Bcl-xL proteins. Moreover, PA-induced expressions of Bcl-2 and Bcl-xL were enhanced by propranolol, an inhibitor of PA phospholydrolase (PAP), whereas completely blocked by mepacrine, an inhibitor of phospholipase A(2) (PLA(2)), suggesting that PLA(2) metabolite of PA is responsible for the increases in Bcl-2 and Bcl-xL protein levels. We further confirmed the involvement of arachidonic acid (AA) in PA-induced survival signals by showing that 1,2-dipalmitoyl-sn-glycero-3-phosphate (DPPA), PA without AA, was unable to increase Bcl-2 and Bcl-xL proteins. Moreover, PA notably increased cyclooxygenase (COX)-2 protein expression, and PA-induced expression of both Bcl-2 and Bcl-xL was inhibited by NS-398, a specific inhibitor of COX-2. Taken together, these findings demonstrate that PA generated by PLD2 plays an important role in cell survival during Fas-mediated apoptosis through the increased Bcl-2 and Bcl-xL protein levels which resulted from PLA(2) and AA-COX2 pathway.  相似文献   

2.
We show that epigallocatechin-3 gallate (EGCG), a major component of green tea, stimulates phospholipase D (PLD) activity in U87 human astroglioma cells. EGCG-induced PLD activation was abolished by the phospholipase C (PLC) inhibitor and a lipase inactive PLC-gamma1 mutant, which is dependent on intracellular or extracellular Ca(2+), with the possible involvement of Ca(2+)/calmodulin-dependent protein kinase II (CaM kinase II). EGCG induced translocation of PLC-gamma1 from the cytosol to the membrane and PLC-gamma1 interaction with PLD1. EGCG regulates the activity of PLD by modulating the redox state of the cells, and antioxidants reverse this effect. Moreover, EGCG-induced PLD activation was reduced by PKC inhibitors or down-regulation of PKC. Taken together, these results show that, in human astroglioma cells, EGCG regulates PLD activity via a signaling pathway involving changes in the redox state that stimulates a PLC-gamma1 [Ins(1,4,5)P(3)-Ca(2+)]-CaM kinase II-PLD pathway and a PLC-gamma1 (diacylglycerol)-PKC-PLD pathway.  相似文献   

3.
Phospholipid signalling is mediated by phospholipid breakdown products generated by phospholipases. The enzymes from animals and plants generating known or potential lipid-derived second messengers are compared. Plants possess a phospholipase C and a phospholipase A2 both of which are agonist-activated. These agonists (auxin, elicitors, perhaps others) bind to the external surface of the plasma membrane. The target enzyme for potential plant lipid-derived second messengers is lipid-activated protein kinase but the possibility that other enzymes may be also lipid-modulated should not be precluded.Abbreviations DAG diacylglycerol - CDPK calmodulin-like domain protein kinase - PLA2 phospholipase A2 - PLC phospholipase C - PLD phospholipase D - PKC protein kinase C - PS phosphatidylserine  相似文献   

4.
Sphingosine-1-phosphate (S1P) is a bioactive sphingolipid capable of regulating critical physiological and pathological functions. Here, we report for the first time that S1P stimulates aldosterone secretion in cells of the zona glomerulosa of the adrenal gland. Regulation of aldosterone secretion is important because this hormone controls electrolyte and fluid balance and is implicated in cardiovascular homeostasis. S1P-stimulated aldosterone secretion was dependent upon the protein kinase C (PKC) isoforms alpha and delta and extracellular Ca2+, and it was inhibited by pertussis toxin (PTX). S1P activated phospholipase D (PLD) through a PTX-sensitive mechanism, also involving PKC alpha and delta and extracellular Ca2+. Primary alcohols, which attenuate the formation of phosphatidic acid (the product of PLD), and cell-permeable ceramides, which inhibit PLD activity, blocked S1P-stimulated aldosterone secretion. Furthermore, propranolol, chlorpromazine, and sphingosine, which are potent inhibitors of phosphatidate phosphohydrolase (PAP) (the enzyme that produces diacylglycerol from phosphatidate), also blocked aldosterone secretion. These data suggest that the PLD/PAP pathway plays a crucial role in the regulation of aldosterone secretion by S1P and that Gi protein-coupled receptors, extracellular Ca2+, and the PKC isoforms alpha and delta are all important components in the cascade of events controlling this process.  相似文献   

5.
Gelsolin, an actin-binding protein, shows a strong ability to bind to phosphatidylinositol 4,5-bisphosphate (PIP(2)). Here we showed in in vitro experiments that gelsolin inhibited recombinant phospholipase D1 (PLD1) and PLD2 activities but not the oleate-dependent PLD and that this inhibition was not reversed by increasing PIP(2) concentration. To investigate the role of gelsolin in agonist-mediated PLD activation, we used NIH 3T3 fibroblasts stably transfected with the cDNA for human cytosolic gelsolin. Gelsolin overexpression suppressed bradykinin-induced activation of phospholipase C (PLC) and PLD. On the other hand, sphingosine 1-phosphate (S1P)-induced PLD activation could not be modified by gelsolin overexpression, whereas PLC activation was suppressed. PLD activation by phorbol myristate acetate or Ca(2+) ionophore A23187 was not affected by gelsolin overexpression. Stimulation of control cells with either bradykinin or S1P caused translocation of protein kinase C (PKC) to the membranes. Translocation of PKC-alpha and PKC-beta1 but not PKC-epsilon was reduced in gelsolin-overexpressed cells, whereas phosphorylation of mitogen-activated protein kinase was not changed. S1P-induced PLC activation and mitogen-activated protein kinase phosphorylation were sensitive to pertussis toxin, but PLD response was insensitive to such treatment, suggesting that S1P induced PLD activation via certain G protein distinct from G(i) for PLC and mitogen-activated protein kinase pathway. Our results suggest that gelsolin modulates bradykinin-mediated PLD activation via suppression of PLC and PKC activities but did not affect S1P-mediated PLD activation.  相似文献   

6.
In the present study, we investigated the involvement of protein kinase C (PKC) in antigen (Ag, DNP-Ascaris suum)-induced phospholipase D (PLD) activation of rat peritoneal mast cells. Phorbor myristate acetate (PMA) as well as Ag activated PLD as inferred by phosphatidylethanol (PEt) production. PKC inhibitors, staurosporine and H-7, however, failed to suppress PMA-stimulated PLD activation, suggesting that PLD activation by PMA is independent of PKC. By contrast, Ag-stimulated PLD activity was significantly reduced by staurosporine and slightly by H-7. Surprisingly, the inhibitors inhibited Ag-stimulated phospholipase C (PLC), correlated to the inhibition of PLD. These observations lead us to conclude that in Ag-stimulated mast cells 1,2-diacylglycerol (DG) formed by PLC directly or indirectly stimulates PLD, independently of PKC.  相似文献   

7.
We have investigated the roles of ceramide in Fas signalling leading to phospholipase D (PLD) activation in A20 cells. Upon stimulation of Fas signalling by anti-Fas monoclonal antibody, sphingomyelin hydrolysis and activation of PLD were induced. Also, the translocation of protein kinase C (PKC) βI and βII and the elevation of diacylglycerol (DAG) content were induced by Fas cross-linking. When phosphatidylcholine-specific phospholipase C (PC-PLC) was inhibited by D609, the Fas-induced changes in PLD activity, DAG content, and PKC translocation were inhibited. In contrast, D609 had no effect on Fas-induced alterations in sphingolipid metabolism, suggesting that changes in ceramide content do not account for Fas-induced PLD activation. Furthermore, C6-ceramide had no effect on Fas-induced PLD activation and PKC translocation. Taken together, these data might suggest that ceramide generated by Fas cross-linking does not affect PKC β-dependent PLD activity stimulated by anti-Fas monoclonal antibody in A20 cells.  相似文献   

8.
Recently, we have isolated a cDNA encoding a muscarinic acetylcholine receptor (mAChR) from Caenorhabditis elegans. To investigate the regulation of phospholipase D (PLD) signaling via a muscarinic receptor, we generated stable transfected Chinese hamster ovary (CHO) cells that overexpress the mAChR of C. elegans (CHO-GAR-3). Carbachol (CCh) induced inositol phosphate formation and a significantly higher Ca(2+) elevation and stimulated PLD activity through the mAChR; this was insensitive to pertussis toxin, but its activity was abolished by the phospholipase C (PLC) inhibitor U73122. Western blot analysis revealed several apparent tyrosine-phosphorylated protein bands after CCh treatment. The CCh-induced PLD activation and tyrosine phosphorylation were significantly reduced by the protein kinase C (PKC) inhibitor calphostin C and down-regulation of PKC and the tyrosine kinase inhibitor genistein. Moreover, the Ca(2+)-calmodulin-dependent protein kinase II (CaM kinase II) inhibitor KN62, in addition to chelation of extracellular or intracellular Ca(2+) by EGTA and BAPTA/AM, abolished CCh-induced PLD activation and protein tyrosine phosphorylation. Taken together, these results suggest that the PLC/PKC-PLD pathway and the CaM kinase II/tyrosine kinase-PLD pathway are involved in the activation of PLD through mAChRs of C. elegans.  相似文献   

9.
In this study we investigated the responses of intracellular calcium ([Ca2+]i) and protein kinase C (PKC) to PDGF in U-1242 MG cells. PDGF-BB stimulated [3H]PDBu binding approximately 2–3 fold. This response was inhibited by preincubating the cells with an inhibitor of phospholipase C (PLC), U73122, suggesting that PLC mediates the induction of PKC translocation by PDGF. PDGF also increased the concentration of [Ca2+]i that was attenuated in a calcium-free medium. This indicates that PDGF-induced elevation of [Ca2+]i is mainly due to influx of extracellular calcium. PDGF-stimulated translocation of PKC was inhibited by the intracellular calcium buffer BAPTA/AM. All gangliosides studied except GM3 inhibited these responses with similar efficacy. Collectively, these results indicate that the signal transduction pathway initiated by PDGF leading to PKC translocation in U-1242 MG cells is intact, and this pathway is inhibited by several gangliosides.Special issue dedicated to Dr. Leon S. Wolfe.  相似文献   

10.
11.
We have previously isolated a 22 kDa protein from a rat brain which was found to be involved in activating phospholipsae D (PLD), and identified the protein as hippocalcin through sequence analysis. Nevertheless, the function of hippocalcin for PLD activation still remains to be resolved. Here, we proposed that hippocalcin was involved in extracellular signal-regulated kinase (ERK)-mediated PLD2 expression. To elucidate a role of hippocalcin, we made hippocalcin transfected NIH3T3 cells and showed that the expression of PLD2 and basal PLD activity were increased in hippocalcin transfected cells. We performed PLD assay with dominant negative PLD2 (DN-PLD2) and hippocalcin co-transfected cells. DN-PLD2 suppressed increase of basal PLD activity in hippocalcin transfected cells, suggesting that increased basal PLD activity is due to PLD2 over-expression. Hippocalcin is a Ca2+-binding protein, which is expressed mainly in the hippocampus. Since it is known that lysophosphatidic acid (LPA) increases intracellular Ca2+, we investigated the possible role of hippocalcin in the LPA-induced elevation of intracellular Ca2+. When the intracellular Ca2+ level was increased by LPA, hippocalcin was translocated to the membrane after LPA treatment in hippocalcin transfected cells. In addition, treatment with LPA in hippocalcin transfected cells markedly potentiated PLD2 expression and showed morphological changes of cell shape suggesting that increased PLD2 expression acts as one of the major factors to cause change of cell shape by making altered membrane lipid composition. Hippocalcin-induced PLD2 expression potentiated by LPA in hippocalcin transfected cells was inhibited by a PI-PLC inhibitor, U73122 and a chelator of intracellular Ca2+, BAPTA-AM suggesting that activation of hippocalcin caused by increased intracellular Ca2+ is important to induce over-expression of PLD2. However, downregulation of PKC and treatment of a chelator of extracellular Ca2+, EGTA had little or no effect on the inhibition of hippocalcin-induced PLD2 expression potentiated by LPA in the hippocalcin transfected cells. Interestingly, when we over-express hippocalcin, ERK was activated, and treatment with LPA in hippocalcin transfected cells significantly potentiated ERK activation. Specific inhibition of ERK dramatically abolished hippocalcin-induced PLD2 expression. Taken together, these results suggest for the first time that hippocalcin can induce PLD2 expression and LPA potentiates hippocalcin-induced PLD2 expression, which is mediated by ERK activation.  相似文献   

12.
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   

13.
Phospholipase D (PLD) is implicated in a variety of physiological processes that reveal it to be a member of the signal transducing phospholipases. We found that PLD1 is activated when basic fibroblast growth factor (bFGF) stimulates neurite outgrowth of an immortalized hippocampal cell line (H19-7). Overexpression of PLD1 in H19-7 cells dramatically elongated bFGF-induced neurite outgrowth and increased PLD activity. Transfection of DN-rPLD1 blocked bFGF-induced PLD activation and completely inhibited neurite outgrowth induced by bFGF, suggesting that PLD1 activation is important in bFGF-induced neurite outgrowth of H19-7 cells. PLD activation and neurite outgrowth induced by bFGF was dependent on phospholipase C gamma (PLC-gamma) and Ca2+, but not protein kinase C (PKC). Furthermore, inhibition of Src and Ras partially blocked bFGF-induced PLD activation and neurite outgrowth, respectively. Coinhibition of Src and Ras completely blocked bFGF-induced PLD activation, suggesting that Src and Ras independently regulate PLD1 activation. Interestingly, bFGF-induced PLD activation and neurite outgrowth did not require ERK1/2 activated by Ras. Taken together, this study demonstrates that bFGF activates PLD1 through PLC-gamma activation, which leads to neurite outgrowth in H19-7 cells. Furthermore, our results show that PLD1 activation by bFGF is regulated by Src and Ras independently.  相似文献   

14.
Inhibition of astrocyte proliferation has been suggested to be an important event in the developmental neurotoxicity associated with ethanol. We have previously shown that the acetylcholine analog carbachol induces astroglial cell proliferation through activation of muscarinic M3 receptors, and that ethanol strongly inhibits this effect by inhibiting activation of protein kinase C (PKC) zeta and its down-stream effector 70-kDa ribosomal S6 kinase (p70S6K). In this study, we investigated whether inhibition by ethanol of this signal transduction pathway in 1321N1 human astrocytoma cells may be due, at least in part, to inhibition of the formation of the PKC zeta activator phosphatidic acid (PA), which is formed by hydrolysis of phosphatidylcholine by phospholipase D (PLD). 1-Butanol, which is a substrate for PLD and inhibits PA formation, inhibited carbachol-induced cell proliferation and the underlying intracellular signaling, whereas its analog tert-butanol, which is a poor substrate for PLD, was much less effective. In addition, exogenous PAs were able to increase DNA synthesis and to activate PKC zeta and p70S6K. Furthermore, in carbachol-stimulated cells, ethanol increased the formation of phosphatidylethanol and inhibited the formation of PA. Taken together, these results indicate that PLD activation plays an important role in carbachol-induced astroglial cell proliferation by generating the second messenger PA, which activates PKC zeta. Moreover, the effect of ethanol on carbachol-induced proliferation appears to be mediated, at least in part, by its ability to interact with PLD leading to a decreased synthesis of PA.  相似文献   

15.
Nitric oxide (NO) and the lipid second messenger phosphatidic acid (PA) are involved in plant defense responses during plant-pathogen interactions. NO has been shown to be involved in the induction of PA production in response to the pathogen associated molecular pattern (PAMP) xylanase in tomato cells. It was shown that NO is critical for PA production induced via phospholipase C (PLC) in concerted action with diacylglycerol kinase (DGK) but not for the xylanase-induced PA via phospholipase D (PLD). In order to study whether this is a general phenomenon during PAMP perception or if it is particular for xylanase, we studied the effect of the PAMP chitosan in tomato cell suspensions. We observed a rapid NO production in tomato cells treated with chitosan. Chitosan induced the formation of PA by activating both PLD and PLC/DGK. The activation of either phospholipase-mediated signaling pathway was inhibited in cells treated with the NO scavenger cPTIO. This indicates that NO is required for PA generation via both the PLD and PLC/DGK pathway during plant defense response in chitosan elicited cells. Responses downstream PA were studied. PLC inhibitors neomycin and U73122 inhibited chitosan-induced ROS production. Differences between xylanase and chitosan-induced phospholipid signaling pathways are discussed.  相似文献   

16.
Many of the effects of 1α,25-(OH)2D3 and 24R,25-(OH)2D3 on costochondral chondrocytes are mediated by the protein kinase C (PKC) signal transduction pathway. 1α,25-(OH)2D3 activates PKC in costochondral growth zone chondrocytes through a specific membrane receptor (1α,25-mVDR), involving rapid increases in diacylglycerol via a phospholipase C (PLC)-dependent mechanism. 24R,25-(OH)2D3 activates PKC in resting zone chondrocytes. Although diacylglycerol is increased by 24R,25-(OH)2D3, PLC is not involved, suggesting a phospholipase D (PLD)-dependent mechanism. Here, we show that resting zone and growth zone cells express mRNAs for PLD1a, PLD1b, and PLD2. Both cell types have PLD activity, but levels are higher in resting zone cells. 24R,25-(OH)2D3, but not 24S,25-(OH)2D3 or 1α,25-(OH)2D3, stimulates PLD activity in resting zone cells within 3 min via nongenomic mechanisms. Neither 1α,25-(OH)2D3 nor 24R,25-(OH)2D3 affected PLD in growth zone cells. Basal and 24R,25-(OH)2D3-stimulated PLD were inhibited by the PLD inhibitors wortmannin and EDS. Inhibition of phosphatidylinositol 3-kinase (PI 3-kinase), PKC, phosphatidylinositol-specific PLC (PI-PLC), and phosphatidylcholine-specific PLC (PC-PLC) had no effect on PLD activity. Thus, 24R,25-(OH)2D3 stimulates PLD, and PI 3-kinase, PI-PLC and PKC are not involved, whereas PLD is required for stimulation of PKC by 24R,25-(OH)2D3. Pertussis toxin, GDPβS, and GTPγS had no effect on 24R,25-(OH)2D3-dependent PLD when added to cell cultures, indicating that G-proteins are not involved. These data show that PKC activation in resting zone cells is mediated by PLD and suggest that a functional 24R,25-(OH)2D3-mVDR is required. The results also support the conclusion that the 24R,25-(OH)2D3-responsive PLD is PLD2, since this PLD isoform is G-protein-independent.  相似文献   

17.
We have investigated the coupling of muscarinic acetylcholine receptors (mAChR) to phospholipid hydrolysis in a human neuroblastoma cell line, LA-N-2, by measuring the formation of 3H-inositol phosphates (3H-IP) and of [3H]phosphatidylethanol ([3H]PEt) in cells prelabeled with [3H]inositol and [3H]oleic acid. The muscarinic agonist carbachol (CCh) stimulated the phospholipase C (PLC)-mediated formation of 3H-IP in a time- and dose-dependent manner (EC50 = 40-55 microM). In addition, in the presence of ethanol (170-300 mM), CCh elevated levels of [3H]PEt [which is regarded as a specific indicator of phospholipase D (PLD) activity] by three- to sixfold. The effect of CCh on PEt formation also was dose dependent (EC50 = 50 microM). Both effects of CCh were antagonized by atropine, indicating that they were mediated by mAChR. Incubation of LA-N-2 cells with the phorbol ester phorbol 12-myristate 13-acetate (PMA, 0.1 microM; 10 min) increased [3H]PEt levels by up to 10-fold. This effect was inhibited by the protein kinase C (PKC) inhibitor staurosporine (1 microM) or by pretreatment for 24 h with 0.1 microM PMA, by 74% and 65%, respectively. In contrast, the effect of CCh on PEt accumulation was attenuated by only 28% in the presence of staurosporine (1 microM). In summary, these results suggest that, in LA-N-2 neuroblastoma cells, mAChR are coupled both to phosphoinositide-specific PLC and to PLD. PKC is capable of stimulating PLD activity in these cells; however, it is not required for stimulation of the enzyme by mAChR activation.  相似文献   

18.
Our previous studies have shown that parathyroid hormone (PTH) stimulates phosphatidylcholine (PC) hydrolysis by phospholipase D (PLD) and transphosphatidylation in UMR-106 osteoblastic cells. To determine whether phospholipase C (PLC) is also involved in the PTH-mediated PC hydrolysis, we used the inhibitor, tricyclodecan-9-yl xanthogenate (D609), a putatively selective antagonist of this pathway. Consistent with this proposed mechanism, D609 decreased (3)H-phosphocholine in extracts from UMR-106 cells prelabeled with (3)H-choline. Unexpectedly, D609 enhanced PC hydrolysis and transphosphatidylation, suggesting that either there was a compensatory increase in PLD activity when PLC was inhibited, or that D609 directly increased PLD activity. The D609-stimulated increase in PC hydrolysis was rapid, being seen as early as 2 min. The effect of D609 was temperature-sensitive, consistent with an enzymatic mechanism. The D609-stimulated increase in PC hydrolysis was PKC-independent, based upon the lack of effect of down-regulation of PKC by phorbol 12,13-dibutyrate on the response. The studies reveal a novel action of this inhibitor on signaling in osteoblastic cells which might influence downstream responses.  相似文献   

19.
The muscarinic agonist carbachol stimulated phospholipase D (PLD) in rat submandibular gland (RSMG) ductal cells in a time and concentration-dependent manner. This effect was inhibited by chelation of extracellular calcium with ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid (EGTA). PLD could also be activated by epinephrine and AlF(4)(-), two polyphosphoinositide-specific phospholipase C (PPI-PLC) activators, and by the phorbol ester o-tetradecanoylphorbol 13-acetate (TPA) which activates protein kinase C (PKC). Ionomycin and thapsigargin only slightly increased PLD activity. Ortho-vanadate, a tyrosine phosphatase inhibitor, also stimulated PLD activity. Both carbachol and o-vanadate increased the formation of inositol phosphates and the tyrosine phosphorylation of at least two proteins (55-60 and 120 kDa). Calphostin C (a PKC inhibitor), U73122 (a PPI-PLC inhibitor) and genistein (a tyrosine kinase inhibitor) blocked the activation of PLD, of PLC and the phosphorylation of tyrosyl residues in response to carbachol and vanadate. Taken together, these results suggest that rat submandibular gland ductal cells express a calcium-dependent PLD activity. This enzyme is regulated by carbachol via a PLC-PKC-tyrosine kinase pathway.  相似文献   

20.
Plant defense mechanisms respond to diverse environmental factors and play key roles in signaling pathways. The phospholipidic signaling pathway forms part of the plant response to several phytoregulators, such as salicylic acid (SA) and methyl jasmonate (MJ), which have been widely used to stimulate secondary metabolite production in cell cultures.1 Furthermore, it has been reported that the levels of such phytoregulators as SA and MJ can increase in response to stressful conditions.2,3 The phospholipidic signal transduction system involves the generation of second messengers by the hydrolysis of phospholipids. In this study, we examined how phospholipidic signaling can be modulated depending on the growth stage of the culture, and we focused on two key lipases having relevant roles in the signaling cascades in plants. An evaluation was made of the effects of SA and MJ on the phospholipase activities in Capsicum chinense Jacq. suspension cells at different phases of the culture cycle. The treatment with SA differentially modified the phospholipase C (PLC) (EC: 3.1.4.3) and phospholipase D (PLD) (EC: 3.1.4.4) activities in a dose-dependent manner that also depended on the day of the culture cycle. In contrast, the treatment with MJ resulted in a biphasic behavior of the PLC and PLD activities. We conclude that the enzymatic activities in the phospholipidic signaling pathways are modified differentially depending on the day of the culture’s growth cycle; accordingly, the response capacity to such environmental factors as phytoregulators is variable at different stages of growth and the physiology of the cells.  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