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 results of studies to evaluate the hypothesis that the 21 kDa GTP-binding protein derived from the ras oncogene is involved in regulation and coupling of hormone receptors to phospholipase activity have thus far been inconsistent. We therefore examined the effect of H-ras transformation on basal, tumor-promoting phorbol ester (TPA)-stimulated, and bradykinin-mediated phospholipid hydrolysis in Madin Darby canine kidney cells (MDCK) by comparing H-ras-transformed MDCK cells (MDCK-RAS) to two non-transformed strains of MDCK cells (MDCK-D1 and MDCK-ATCC). In unstimulated MDCK-RAS, diacylglycerol (DAG), inositol phosphate accumulation, and choline phosphate release were increased while arachidonic acid and arachidonic acid metabolite (AA) release was not increased, suggesting that ras transformation increased phospholipase C activity. Protein kinase C (PK-C) activity was decreased, and specific binding of [3H]phorbol ester was reduced in MDCK-RAS relative to the non-transformed MDCK cells suggesting that elevated DAG may activate and thereby down-regulate PK-C. Consistent with this finding in MDCK-RAS, TPA-stimulated AA release and subsequent prostaglandin E2 production were decreased, while TPA-stimulated choline phosphate release was increased. Bradykinin receptor-stimulated phospholipid hydrolysis in MDCK-RAS was similar to that of non-transformed cells, suggesting that the ras-derived protein does not directly couple bradykinin receptors to phospholipases in MDCK cells. However, the ability of TPA-treatment to inhibit bradykinin-stimulated phosphoinositide hydrolysis and enhance bradykinin-stimulated AA release was attenuated in MDCK-RAS. Additionally, in MDCK-RAS the conversion of arachidonic acid to prostaglandin E2 was substantially reduced. We conclude that ras transformation of MDCK cells increases DAG levels, thereby activating and, in turn, down-regulating PK-C and certain responses to TPA. Since activation of PK-C may result in a variety of effects on signal transduction pathways, we propose that increased DAG and altered PK-C levels associated with ras transformation may account for the inconsistent effects previously observed in studies evaluating the effect of ras transformation on phospholipases and other signal transduction systems.  相似文献   

2.
alpha 1-Adrenergic receptors mediate two effects on phospholipid metabolism in Madin-Darby canine kidney (MDCK-D1) cells: hydrolysis of phosphoinositides and arachidonic acid release with generation of prostaglandin E2 (PGE2). The similarity in concentration dependence for the agonist (-)-epinephrine in eliciting these two responses implies that they are mediated by a single population of alpha 1-adrenergic receptors. However, we find that the kinetics of the two responses are quite different, PGE2 production occurring more rapidly and transiently than the hydrolysis of phosphoinositides. The antibiotic neomycin selectively decreases alpha 1-receptor-mediated phosphatidylinositol 4,5-bisphosphate hydrolysis without decreasing alpha 1-receptor-mediated arachidonic acid release and PGE2 generation. In addition, receptor-mediated inositol trisphosphate formation is independent of extracellular calcium, whereas release of labeled arachidonic acid is largely calcium-dependent. Moreover, based on studies obtained with labeled arachidonic acid, receptor-mediated generation of arachidonic acid cannot be accounted for by breakdown of phosphatidylinositol monophosphate, phosphatidylinositol bisphosphate, or phosphatidic acid. Further studies indicate that epinephrine produces changes in formation or turnover of several classes of membrane phospholipids in MDCK cells. We conclude that alpha 1-adrenergic receptors in MDCK cells appear to regulate phospholipid metabolism by the parallel activation of phospholipase C and phospholipase A2. This parallel activation of phospholipases contrasts with models described in other systems which imply sequential activation of phospholipase C and diacylglycerol lipase or phospholipase A2.  相似文献   

3.
Vasopressin and oxytocin both stimulated inositol phosphate accumulation in isolated uterine decidua cells. Pretreatment of cells with the phorbol ester 12-O-tetradecanoylphorbol 13-acetate (TPA) prevented this agonist-induced phosphoinositide hydrolysis. TPA (0.1 microM) alone had no effect on basal inositol phosphate accumulation, but stimulated phosphoinositide deacylation, as indicated by a 2-fold increase in lysophosphatidylinositol and glycerophosphoinositol. TPA also stimulated a dose-related release of arachidonic acid from decidua-cell phospholipid [phosphatidylcholine (PC) much greater than phosphatidylinositol (PI) greater than phosphatidylethanolamine]. The phorbol ester 4 beta-phorbol 12,13-diacetate (PDA) at 0.1 microM had no effect on arachidonic acid mobilization. The TPA-stimulated increase in arachidonic acid release was apparent by 2 1/2 min (116% of control), maximal after 20 min (283% of control), and remained around this value (306% of control) after 120 min incubation. TPA also stimulated significant increases in 1,2-diacylglycerol and monoacylglycerol production at 20 and 120 min. Although the temporal increases in arachidonic acid and monoacylglycerol accumulation in the presence of TPA continued up to 120 min, that of 1,2-diacylglycerol declined after 20 min. In decidua cells prelabelled with [3H]choline, TPA also stimulated a significant decrease in radiolabelled PC after 20 min, which was accompanied by an increased release of water-soluble metabolites into the medium. Most of the radioactivity in the extracellular pool was associated with choline, whereas the main cellular water-soluble metabolite was phosphorylcholine. TPA stimulated extracellular choline accumulation to 183% and 351% of basal release after 5 and 20 min respectively and cellular phosphorylcholine production to 136% of basal values after 20 min. These results are consistent with a model in which protein kinase C activation by TPA leads to arachidonic acid mobilization from decidua-cell phospholipid by a mechanism involving phospholipase A-mediated PI hydrolysis and phospholipase C-mediated PC hydrolysis, coupled with further hydrolysis of the 1,2-diacylglycerol product.  相似文献   

4.
Bradykinin (BK) or kallikreins activate B2 receptors (R) that couple Galpha(i) and Galpha(q) proteins to release arachidonic acid (AA) and elevate intracellular Ca2+ concentration ([Ca2+]i). Thrombin cleaves the protease-activated-receptor-1 (PAR1) that couples Galpha(i), Galpha(q), and Galpha(12/13) proteins. In Chinese hamster ovary cells stably transfected with human B2R, thrombin liberated little AA, but it significantly potentiated AA release by B2R agonists. We explored mechanisms of cooperativity between constitutively expressed PAR1 and B2R. We also examined human endothelial cells expressing both Rs constitutively. The PAR1 agonist hexapeptide (TRAP) was as effective as thrombin. Inhibitors of components of Galpha(i), Galpha(q), and Galpha(12/13) signaling pathways, and a protein kinase C (PKC)-alpha inhibitor, G?-6976, blocked potentiation, while phorbol, an activator, enhanced it. Several inhibitors, including a RhoA kinase inhibitor, a [Ca2+]i antagonist, and an inositol-(1,3,4)-trisphosphate R antagonist, reduced mobilization of [Ca2+]i by thrombin and blocked potentiation of AA release by B2R agonists. Because either a nonselective inhibitor (isotetrandrine) of phospholipase A2 (PLA2) or a Ca2+-dependent PLA2 inhibitor abolished potentiation of AA release by thrombin, while a Ca2+-independent PLA2 inhibitor did not, we concluded that the mechanism involves Ca2+-dependent PLA2 activation. Both thrombin and TRAP modified activation and phosphorylation of the B2R induced by BK. In lower concentrations they enhanced it, while higher concentrations inhibited phosphorylation and diminished B2R activation. Protection of the NH2-terminal Ser1-Phe2 bond of TRAP by an aminopeptidase inhibitor made this peptide much more active than the unprotected agonist. Thus PAR1 activation enhances AA release by B2R agonists through signal transduction pathway.  相似文献   

5.
Acute hydrolysis of phosphoinositides has been demonstrated in bovine aortic endothelial cells (BAEC) treated with bradykinin (BK) (10(-7)M). The first phosphoinositide to decrease was phosphatidylinositol-4,5-bisphosphate (PIP2) indicating this to be the initial substrate of phospholipase action. Other lipid changes associated with the stimulation of BAEC were an increase in diacylglycerol (DAG) and arachidonic acid (AA) with a sustained production of phosphatidic acid (PA). The changes in cell phospholipids were accompanied by the release of inositol phosphates. Inositol-1,4,5-trisphosphate (Ins-1,4,5-P3) was produced within 10 s of stimulation with BK. There was no evidence for the production of inositol-1,3,4-trisphosphate. The release of ionic calcium (Ca2+) intracellularly was demonstrated. The timecourse of the rise in intracellular Ca2+ was consistent with the timecourse of production of IP3. Intracellular Ca2+ rose from 127 +/- 21 nM to 462 +/- 27 nM. The Ca2+ peak was at 7.0 +/- 0.4 s and took 3 min to reach a steady state which remained above the basal level. When extracellular Ca2+ was depleted in the extracellular medium a spike of intracellular Ca2+ release was measured with an immediate return to basal. Entry of extracellular Ca2+ into the cell after ionophore A23187 treatment does not induce inositol phosphate release, indicating that phosphoinositide hydrolysis is likely to be the cause rather than consequence of the elevation in cytosolic Ca2+. These data indicate action of phospholipase C (PLC) on PIP2 after BK stimulation of BAEC with the subsequent production of InsP3 causing the resulting intracellular Ca2+ release.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Regulatory mechanisms in bradykinin (BK) activated release of arachidonate (ARA) and synthesis of prostaglandin (PG) and platelet activating factor (PAF) were studied in bovine pulmonary artery endothelial cells (BPAEC). A role for GTP binding protein (G-protein) in the binding of BK to the cells was determined. Guanosine 5-O- (thiotriphosphate), (GTPtauS), lowered the binding affinity for BK and increased the Kd for the binding from 0.45 to 1.99 nM. The Bmax remained unaltered at 2.25 x 10(-11) mole. Exposure of the cells to aluminium fluoride also reduced the affinity for BK. Bradykinin-induced release of ARA proved pertussis toxin (PTX) sensitive, with a maximum sensitivity at 10 ug/ml PTX. GTPtauS at 100 muM increased the release of arachidonate. The effect of GTPtauS and BK was additive at suboptimal doses of BK up to 0.5 nM but never exceeded the levels of maximal BK stimulation at 50 nM. PTX also inhibited the release of ARA induced by the calcium ionophore, A23187. Phorbol 12-myristate 13-acetate or more commonly known as tetradecanoyl phorbol acetate (TPA) itself had little effect on release by the intact cells. However, at 100 nM it augmented the BK activated release. This was downregulated by overnight exposure to TPA and correlated with down-regulation of protein kinase C (PKC) activity. The down-regulation only affected the augmentation of ARA release by TPA but not the original BK activated release. TPA displayed a similar, but more potent amplification of PAF synthesis in response to both BK or the calcium ionophore A23187. These results taken together point to the participation of G-protein in the binding of BK to BPAEC and its activation of ARA release. Possibly two types of G-protein are involved, one associated with the receptor, the other activated by Ca(2+) and perhaps associated with phospholipase A(2) (PLA(2)). Our results further suggest that a separate route of activation, probably also PLA(2) related, takes place through a PKC catalysed phosphorylation.  相似文献   

7.
In cloned osteoblast-like cells, MC3T3-E1, prostaglandin F2 alpha (PGF2 alpha) stimulated arachidonic acid (AA) release in a dose-dependent manner in the range between 1 nM and 10 microM. 12-O-tetradecanoylphorbol-13-acetate (TPA), a protein kinase C (PKC) activator, which by itself had little effect on AA release, markedly amplified the release of AA stimulated by PGF2 alpha in a dose-dependent manner. 4 alpha-phorbol 12,13-didecanoate, a phorbol ester which is inactive for PKC, showed little effect on the PGF2 alpha-induced AA release. 1-oleoyl-2-acetylglycerol (OAG), a specific activator for PKC, mimicked TPA by enhancement of the AA release induced by PGF2 alpha. H-7, a PKC inhibitor, markedly suppressed the effect of OAG on PGF2 alpha-induced AA release. Quinacrine, a phospholipase A2 inhibitor, showed partial inhibitory effect on PGF2 alpha-induced AA release, while it suppressed the amplification by OAG of PGF2 alpha-induced AA release almost to the control level. Furthermore, TPA enhanced the AA release induced by melittin, known as a phospholipase A2 activator. On the other hand, TPA inhibited the formation of inositol trisphosphate stimulated by PGF2 alpha. Under the same condition, PGF2 alpha indeed stimulated prostaglandin E2 (PGE2) synthesis and TPA markedly amplified the PGF2 alpha-induced PGE2 synthesis as well as AA release. These results indicate that the activation of PKC amplifies PGF2 alpha-induced both AA release and PGE2 synthesis through the potentiation of phospholipase A2 activity in osteoblast-like cells.  相似文献   

8.
Earlier studies have shown that bradykinin stimulated release of catecholamines from chromaffin cells by an influx of calcium through dihydropyridine-insensitive channels, and also that bradykinin stimulated (poly)phosphoinositide hydrolysis. To investigate membrane-bound second messengers in chromaffin cells, and to elucidate any role these may play in stimulus-secretion coupling, we have studied the influence of bradykinin on diacylglycerol and phosphatidic acid (PA). Using equilibrium labelling of primary cultures of chromaffin cells with [3H]arachidonic acid or [3H]glycerol, we found no influence of bradykinin (10 nM) on labelled diacylglycerol formation, either in the presence or absence of inhibitors of diacylglycerol lipase or kinase. However, when we used cells prelabelled with 32Pi for 2.5 h, we found that bradykinin produced a substantial stimulation of label found in PA, with an EC50 value of about 1 nM. This bradykinin stimulation of [32P]PA formation was only partially dependent on extracellular calcium, in contrast to the smaller response to nicotine, which was completely dependent on extracellular calcium. Short (10 min) pretreatment with tetradecanoylphorbol acetate (TPA) almost completely eliminated the bradykinin-stimulated formation of inositol phosphates, but failed to affect bradykinin stimulation of label in PA, suggesting that PA production in response to bradykinin is not downstream of phospholipase C activation. TPA alone failed to stimulate [32P]PA substantially, whereas long-term (24 or 48 h) treatment with TPA failed to attenuate the response to bradykinin. Diacylglycerol kinase inhibitors were also without effect on the bradykinin stimulation of [32P]PA. These results suggest that bradykinin stimulates PA production by a mechanism independent of the activation of protein kinase C.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Kallikreins cleave plasma kininogens to release the bioactive peptides bradykinin (BK) or kallidin (Lys-BK). These peptides then activate widely disseminated B2 receptors with consequences that may be either noxious or beneficial. We used cultured cells to show that kallikrein can bypass kinin release to activate BK B2 receptors directly. To exclude intermediate kinin release or kininogen uptake from the cultured medium, we cultured and maintained cells in medium entirely free of animal proteins. We compared the responses of stably transfected Chinese hamster ovary (CHO) cells that express human B2 receptors (CHO B2) and cells that coexpress angiotensin I-converting enzyme (ACE) as well (CHO AB). We found that BK (1 nM or more) and tissue kallikrein (1-10 nM) both significantly increased release of arachidonic acid beyond unstimulated baseline level. An enzyme-linked immunoassay for kinin established that kallikrein did not release a kinin from CHO cells. We confirmed the absence of kininogen mRNA with RT-PCR to rule out kininogen synthesis by CHO cells. We next tested an ACE inhibitor for enhanced BK receptor activation in the absence of kinin release and synthesized an ACE-resistant BK analog as a control for these experiments. Enalaprilat (1 microM) potentiated kallikrein (100 nM) in CHO AB cells but was ineffective in CHO B2 cells that do not bear ACE. We concluded that kallikrein activated B2 receptors without releasing a kinin. Furthermore, inhibition of ACE enhanced the receptor activation by kallikrein, an action that may contribute to the manifold therapeutic effects of ACE inhibitors.  相似文献   

10.
Phospholipid remodeling resulting in arachidonic acid (AA) release and metabolism in human neutrophils stimulated by calcium ionophore A23187 has been extensively studied, while data obtained using physiologically relevant stimuli is limited. Opsonized zymosan and immune complexes induced stimulus-specific alterations in lipid metabolism that were different from those induced by A23187. [3H]AA release correlated with activation of phospholipase A2 (PLA2) but not with cellular activation as indicated by superoxide generation. The latter correlated more with calcium-dependent phospholipase C (PLC) activation and elevation of cellular diacylglycerol (DAG) levels. When cells that had been allowed to incorporate [3H]AA were stimulated with A23187, large amounts of labeled AA was released, most of which was metabolized to 5-HETE and leukotriene B4. Stimulation with immune complexes also resulted in the release of [3H]AA but this released radiolabeled AA was not metabolized. In contrast, stimulation with opsonized zymosan induced no detectable release of [3H]AA. Analysis of [3H]AA-labeled lipids in resting cells indicated that the greatest amount of label was incorporated into the phosphatidylinositol (PI) pool, followed closely by phosphatidylcholine and phosphatidylserine, while little [3H]AA was detected in the phosphatidylethanolamine pool. During stimulation with A23187, a significant decrease in labeled PI occurred and labeled free fatty acid in the pellet increased. With immune complexes, only a small decrease was seen in labeled PI while the free fatty acid in the pellets was unchanged. In contrast, opsonized zymosan decreased labeled PI, and increased labeled DAG. Phospholipase activity in homogenates from human neutrophils was also assayed. A23187 and immune complexes, but not zymosan, significantly enhanced PLA2 activity in the cell homogenates. On the other hand, PLC activity was enhanced by zymosan and immune complexes. Stimulated increases in PLC activity correlated with enhanced superoxide generation induced by the stimulus.  相似文献   

11.
The effect of nerve growth factor on the metabolism of arachidonic acid and the hydrolysis of phosphatidylinositol in PC12 cells was examined. Addition of nerve growth factor to PC12 cells isotopically labeled with [3H]arachidonic acid caused an increased release of radioactivity. In a similar manner, treatment of PC12 cells prelabeled with [3H]inositol increased inositol monophosphate accumulation in the presence of LiCl. Stimulation of [3H]arachidonic acid release by nerve growth factor was concentration dependent, attaining a maximum at 0.5 nM. Concentrations of nerve growth factor above 0.5 nM caused less than maximal stimulation. In contrast, nerve growth factor-stimulated accumulation of [3H]inositol monophosphate exhibited a sigmoidal dose-response curve with an apparent maximum at 8 nM. Increased accumulation of [3H]inositol monophosphate could be detected as early as 60 s after nerve growth factor addition, whereas nerve growth factor-stimulated release of [3H]arachidonic acid was not observed until 5 min after nerve growth factor treatment. The nerve growth factor-stimulated release of [3H]arachidonic acid was independent of extracellular calcium concentration. Increased [3H]inositol monophosphate accumulation elicited by nerve growth factor was dependent on the presence of extracellular calcium. These results suggest that the increased metabolism of arachidonic acid and the enhanced hydrolysis of phosphatidylinositol are separately regulated by nerve growth factor.  相似文献   

12.
The kinin peptides are released during inflammation and are amongst the most potent known mediators of vasodilatation, pain, and oedema. A role in the modulation or induction of healthy breast tissue growth has been postulated for tissue kallikrein present in human milk. Moreover, tissue kallikrein was found in malignant human breast tissue and bradykinin (BK) stimulates the proliferation of immortalised breast cancer cells. Aim of the present article was to investigate whether BK also exerts mitogenic activity in normal breast epithelial cells and partially characterise the signalling machinery involved. Results show that BK increased up to 2-fold the 24 h proliferation of breast epithelial cells in primary culture, and that the BK B2 receptor (not B1) inhibitor alone fully blocked the BK response. Intracellular effects of B2 stimulation were the following: (a) the increase of free intracellular Ca(2+) concentration by a mechanism dependent upon the phospholipase C (PLC) activity; (b) the cytosol-to-membrane translocation of conventional (PKC)-alpha and -beta isozymes, novel PKC-delta, -epsilon, and -eta isozymes; (c) the phosphorylation of the extracellular-regulated kinase 1 and 2 (ERK1/2); and (d) the stimulation of the expression of c-Fos protein. EGF, a well known stimulator of cell proliferation, regulated the proliferative response in human epithelial breast cells to the same extent of BK. The effects of BK on proliferation, ERK1/2 phosphorylation, and c-Fos expression were abolished by GF109203X, which inhibits PKC-delta isozyme. Conversely, G?6976, an inhibitor of PKC-alpha and -beta isozymes, and the 18-h treatment of cells with PMA, that led to the complete down-regulation of PKC-alpha, -beta, -epsilon, and -eta, but not of PKC-delta, did not have any effect, thereby indicating that the PKC-delta mediates the mitogenic signalling of BK. Phosphoinositide 3-kinase (PI3K), tyrosine kinase of the epidermal growth factor receptor (EGFR), and mitogen activated protein kinase kinases (MEK) inhibitors were also tested. The results suggest that EGFR, PI3K, and ERK are required for the proliferative effects of BK. In addition, the BK induced cytosol-to-membrane translocation of PKC-delta was blocked by PI3K inhibition, suggesting that PI3K is upstream to PKC-delta. In conclusion, BK has mitogenic actions in cultured human epithelial breast cells; the activation of PKC-delta through B2 receptor acts in concert with ERK and PI3K pathways to induce cell proliferation.  相似文献   

13.
Previous studies suggested a role for calcium in CYP2E1-dependent toxicity. The possible role of phospholipase A2 (PLA2) activation in this toxicity was investigated. HepG2 cells that overexpress CYP2E1 (E47 cells) exposed to arachidonic acid (AA) +Fe-NTA showed higher toxicity than control HepG2 cells not expressing CYP2E1 (C34 cells). This toxicity was inhibited by the PLA2 inhibitors aristolochic acid, quinacrine, and PTK. PLA2 activity assessed by release of preloaded [3H]AA after treatment with AA+Fe was higher in the CYP2E1 expressing HepG2 cells. This [3H]AA release was inhibited by PLA2 inhibitors, alpha-tocopherol, and by depleting Ca2+ from the cells (intracellular + extracellular sources), but not by removal of extracellular calcium alone. Toxicity was preceded by an increase in intracellular calcium caused by influx from the extracellular space, and this was prevented by PLA2 inhibitors. PLA2 inhibitors also blocked mitochondrial damage in the CYP2E1-expressing HepG2 cells exposed to AA+Fe. Ca2+ depletion and removal of extracellular calcium inhibited toxicity at early time periods, although a delayed toxicity was evident at later times in Ca2+-free medium. This later toxicity was also inhibited by PLA2 inhibitors. Analogous to PLA2 activity, Ca2+ depletion but not removal of extracellular calcium alone prevented the activation of calpain activity by AA+Fe. These results suggest that release of stored calcium by AA+Fe, induced by lipid peroxidation, can initially activate calpain and PLA2 activity, that PLA2 activation is critical for a subsequent increased influx of extracellular Ca2+, and that the combination of increased PLA2 and calpain activity, increased calcium and oxidative stress cause mitochondrial damage, that ultimately produces the rapid toxicity of AA+Fe in CYP2E1-expressing HepG2 cells.  相似文献   

14.
The effects of the neuropeptide bradykinin (BK) and its natural proteolytic fragment Des-Arg9 bradykinin (DBK) on DNA synthesis and phospholipase C activation were investigated in cultured mesangial cells. DBK, acting through a distinct bradykinin receptor, induced DNA synthesis in serum-starved cultured mesangial cells. The effect of DBK was dose dependent (ED50 = 0.6 microM) and was strongly potentiated by insulin. Under the same conditions, BK had no effect. Down-regulation of protein kinase C by long term pretreatment with 12-O-tetradecanoylphorbol-13-acetate (TPA) markedly reduced DBK-induced DNA synthesis. In the same way, co-incubation with the protein kinase C inhibitor staurosporine potently attenuated the response to DBK, suggesting a role of protein kinase C in DBK-induced mitogenesis. Analysis of phosphoproteins from 32P-labeled mesangial cells by two-dimensional gel electrophoresis revealed that DBK, like TPA but not BK, induced a net increase in the phosphorylation of an acidic cellular protein migrating with an apparent Mr = 80,000 (termed 80K), identified as a major and specific substrate of protein kinase C. Phosphorylation of the 80K protein by DBK or TPA was completely abolished in cells depleted of protein kinase C. DBK and TPA also induced an increase in phosphorylation of an Mr = 28,000 protein. Moreover, DBK but not TPA stimulated the phosphorylation of an Mr = 18,000 protein in normal as well as in protein kinase C-depleted cells. Analysis of phospholipase C activation revealed that DBK induced a large and sustained increase in diacylglycerol production and inositol phosphate accumulation over a 10-min incubation. BK had only a minor effect on both parameters. These results demonstrate that DBK, but not BK, modulates DNA synthesis through protein kinase C activation in cultured mesangial cells.  相似文献   

15.
The tumor promoter 12-O-tetradecanoylphorbol 13-acetate (TPA) induces release of arachidonic acid (AA) from HeLa cells with a maximum at 2-3 h. Subsequently the extracellular level of AA decreases. Cycloheximide (CH, 10(-5) M does not influence the release of AA, however, it causes the AA level to remain elevated. In the presence of TPA and CH (i) re-uptake of AA is not altered, (ii) re-incorporation of AA into phosphatidylinositol (and phosphatidylethanolamine) is largely increased, and (iii) the level of lysophosphatidylinositol is elevated. The latter two phenomena can be prevented by fluocinolone acetonide (10(-8) M), i.e. by inhibition of phospholipase A2 (PLA2). These data point to a continuously elevated PLA2 activity in the presence of TPA and CH. The phorbol ester appears to induce a proteinaceous principle which diminishes PLA2 activity.  相似文献   

16.
Murine bone marrow-derived macrophages were induced to prostaglandin synthesis by activators of protein kinase C, the phorbolester TPA and the diacylglycerols dioctanoylglycerol (diC8) and diolein (diC18:1). As short term stimulation of prostaglandin synthesis is mainly dependent on the availability of free arachidonic acid, the modulation of arachidonic acid liberation and reacylation was investigated. DiC8 inhibited the reacylating enzyme lysophosphatide acyltransferase in the in vitro assay, but there was no evidence for an inhibitory effect of TPA or diacylglycerols on the activity of the lysophosphatide acyltransferase in whole cells. The release of arachidonic acid from prelabelled cells was stimulated by TPA and the diacylglycerols even in the presence of an inhibitor of reacylation, indicating an activation of phospholipase A2. An activation of phospholipase A2 was measured in membranes derived from TPA-stimulated macrophages. These data indicate that the enhanced pool of free arachidonic acid, which drives prostaglandin synthesis, is primarily due to a stimulation of the liberation of arachidonic acid from membrane phospholipids.  相似文献   

17.
Activation of group IV cytosolic phospholipase A(2) (gIV-PLA(2)) is the essential first step in the synthesis of inflammatory eicosanoids and in integrin-mediated adhesion of leukocytes. Prior investigations have demonstrated that phosphorylation of gIV-PLA(2) results from activation of at least two isoforms of mitogen-activated protein kinase (MAPK). We investigated the potential role of phosphoinositide 3-kinase (PI3K) in the activation of gIV-PLA(2) and the hydrolysis of membrane phosphatidylcholine in fMLP-stimulated human blood eosinophils. Transduction into eosinophils of Deltap85, a dominant negative form of class IA PI3K adaptor subunit, fused to an HIV-TAT protein transduction domain (TAT-Deltap85) concentration dependently inhibited fMLP-stimulated phosphorylation of protein kinase B, a downstream target of PI3K. FMLP caused increased arachidonic acid (AA) release and secretion of leukotriene C(4) (LTC(4)). TAT-Deltap85 and LY294002, a PI3K inhibitor, blocked the phosphorylation of gIV-PLA(2) at Ser(505) caused by fMLP, thus inhibiting gIV-PLA(2) hydrolysis and production of AA and LTC(4) in eosinophils. FMLP also caused extracellular signal-related kinases 1 and 2 and p38 MAPK phosphorylation in eosinophils; however, neither phosphorylation of extracellular signal-related kinases 1 and 2 nor p38 was inhibited by TAT-Deltap85 or LY294002. Inhibition of 1) p70 S6 kinase by rapamycin, 2) protein kinase B by Akt inhibitor, or 3) protein kinase C by Ro-31-8220, the potential downstream targets of PI3K for activation of gIV-PLA(2), had no effect on AA release or LTC(4) secretion caused by fMLP. We find that PI3K is required for gIV-PLA(2) activation and hydrolytic production of AA in activated eosinophils. Our data suggest that this essential PI3K independently activates gIV-PLA(2) through a pathway that does not involve MAPK.  相似文献   

18.
Inbred Fischer 344 rats display airway hyperresponsiveness (AHR) in vivo compared with the normoresponsive Lewis strain. Fischer AHR has been linked with increased airway smooth muscle (ASM) contraction ex vivo and enhanced ASM cell intracellular Ca(2+) mobilization in response to serotonin compared with Lewis. To determine the generality of this association, we tested whether bradykinin (BK) also stimulates greater contraction of Fischer airways and greater Ca(2+) mobilization in Fischer ASM cells. Explants of Fischer intraparenchymal airways constricted faster and to a greater degree in response to BK than Lewis airways. BK also evoked higher Ca(2+) transients in Fischer than in Lewis ASM cells. ASM cell B(2) receptor expression was similar between the two strains. BK activated both phosphatidylinositide-specific phospholipase C (PI-PLC) and phosphatidylcholine-specific PLC to mobilize Ca(2+) in Fischer and Lewis ASM cells. PI-PLC activity, as measured by inositol polyphosphate accumulation, was similar in the two strains. PKC inhibition with GF109203X, Go6973, or Go6983 attenuated BK-mediated Ca(2+) transients in Fischer cells, whereas GF109203X potentiated while Go6976 and Go6983 did not affect Ca(2+) transients in Lewis cells. Enhanced Ca(2+) mobilization in ASM cells can arise from variations in PKC and may be an important component of nonspecific, innate AHR.  相似文献   

19.
20.
The pharmacological properties of bradykinin (BK) receptors were characterized in canine cultured corneal epithelial cells (CECs) using [(3)H]-BK as a radioligand. Analysis of binding isotherms gave an apparent equilibrium dissociation constant of 0.34 +/- 0.07 nM and a maximum receptor density of 179 +/- 23 fmol/mg protein. Neither a B(1) receptor-selective agonist (des-Arg(9)-BK) nor antagonist ([Leu(8), des-Arg(9)]-BK) significantly inhibited [(3)H]-BK binding to CECs, thus excluding the presence of B(1) receptors in canine CECs. The specific binding of [(3)H]-BK to CECs was inhibited by B(2) receptor-selective agonists (BK and kallidin) and antagonists (Hoe 140 and [D-Arg(0), Hyp(3), Thi(5,8), D-Phe(7)]-BK), with a best fit using a one-binding-site model. The order of potency for the inhibition of [(3)H]-BK binding was BK = Hoe 140 > kallidin > [D-Arg(0), Hyp(3), Thi(5,8), D-Phe(7)]-BK. Stimulation of CECs by BK produced a concentration-dependent accumulation of inositol phosphates (IP) and an initial transient peak of intracellular Ca(2+). B(2) receptor-selective antagonist ([D-Arg(0), Hyp(3), Thi(5,8), D-Phe(7)]-BK) significantly antagonized the BK-induced responses with dissociation constants of 6.0-6.1. Pretreatment of CECs with pertussis toxin (PTX) or cholera toxin did not alter the BK-induced IP accumulation. Incubation of CECs in the absence of external Ca(2+) led to a significant attenuation of the IP accumulation induced by BK. These results demonstrate that BK directly stimulates phospholipase C-mediated signal transduction through BK B(2) receptors via a PTX-insensitive G protein in canine CECs. This effect may function as the transducing mechanism for BK-mediated cellular responses.  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