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.
2.
Murine P388D(1) macrophages exhibit a delayed prostaglandin biosynthetic response when exposed to bacterial lipopolysaccharide (LPS) for prolonged periods of time that is dependent on induction of the genes coding for Group V secretory phospholipase A(2) and cyclooxygenase-2. We herein report that LPS-induced arachidonic acid (AA) metabolite release in P388D(1) macrophages is strongly attenuated by the P2X(7) purinergic receptor antagonists periodate-oxidized ATP and pyridoxal-phosphate-6-azophenyl-2', 4'-disulfonic acid, and this is accompanied by suppression of the expression of both Group V secretory phospholipase A(2) and cyclooxygenase-2. The effect appears to be specific for LPS, because the P2 purinergic receptor antagonists do not affect P388D(1) cell stimulation by other stimuli such as platelet-activating factor or the Ca(2+) ionophore A23187. Moreover, extracellular nucleotides are found to stimulate macrophage AA mobilization with a pharmacological profile that implicates the participation of the P2X(7) receptor and that is inhibited by periodate-oxidized ATP. Collectively these results demonstrate coupling of the P2X(7) receptor to the AA cascade in P388D(1) macrophages and implicate the participation of this type of receptor in LPS-induced AA mobilization.  相似文献   

3.
LPS is a potent stimulator of bone resorption in inflammatory diseases. The mechanism by which LPS induces osteoclastogenesis was studied in cocultures of mouse osteoblasts and bone marrow cells. LPS stimulated osteoclast formation and PGE(2) production in cocultures of mouse osteoblasts and bone marrow cells, and the stimulation was completely inhibited by NS398, a cyclooxygenase-2 inhibitor. Osteoblasts, but not bone marrow cells, produced PGE(2) in response to LPS. LPS-induced osteoclast formation was also inhibited by osteoprotegerin (OPG), a decoy receptor of receptor activator of NF-kappaB ligand (RANKL), but not by anti-mouse TNFR1 Ab or IL-1 receptor antagonist. LPS induced both stimulation of RANKL mRNA expression and inhibition of OPG mRNA expression in osteoblasts. NS398 blocked LPS-induced down-regulation of OPG mRNA expression, but not LPS-induced up-regulation of RANKL mRNA expression, suggesting that down-regulation of OPG expression by PGE(2) is involved in LPS-induced osteoclast formation in the cocultures. NS398 failed to inhibit LPS-induced osteoclastogenesis in cocultures containing OPG knockout mouse-derived osteoblasts. IL-1 also stimulated PGE(2) production in osteoblasts and osteoclast formation in the cocultures, and the stimulation was inhibited by NS398. As seen with LPS, NS398 failed to inhibit IL-1-induced osteoclast formation in cocultures with OPG-deficient osteoblasts. These results suggest that IL-1 as well as LPS stimulates osteoclastogenesis through two parallel events: direct enhancement of RANKL expression and suppression of OPG expression, which is mediated by PGE(2) production.  相似文献   

4.
In Madin-Darby canine kidney D1 cells extracellular nucleotides activate P2Y receptors that couple to several signal transduction pathways, including stimulation of multiple phospholipases and adenylyl cyclase. For one class of P2Y receptors, P2Y2 receptors, this stimulation of adenylyl cyclase and increase in cAMP occurs via the conversion of phospholipase A2 (PLA2)-generated arachidonic acid (AA) to prostaglandins (e.g. PGE2). These prostaglandins then stimulate adenylyl cyclase activity, presumably via activation of prostanoid receptors. In the current study we show that agents that increase cellular cAMP levels (including PGE2, forskolin, and the beta-adrenergic agonist isoproterenol) can inhibit P2Y receptor-promoted AA release. The protein kinase A (PKA) inhibitor H89 blocks this effect, suggesting that this feedback inhibition occurs via activation of PKA. Studies with PGE2 indicate that inhibition of AA release is attributable to inhibition of mitogen-activated protein kinase activity and in turn of P2Y receptor stimulated PLA2 activity. Although cAMP/PKA-mediated inhibition occurs for P2Y receptor-promoted AA release, we did not find such inhibition for epinephrine (alpha1-adrenergic) or bradykinin-mediated AA release. Taken together, these results indicate that negative feedback regulation via cAMP/PKA-mediated inhibition of mitogen-activated protein kinase occurs for some, but not all, classes of receptors that promote PLA2 activation and AA release. We speculate that receptor-selective feedback inhibition occurs because PLA2 activation by different receptors in Madin-Darby canine kidney D1 cells involves the utilization of different signaling components that are differentially sensitive to increases in cAMP or, alternatively, because of compartmentation of signaling components.  相似文献   

5.
P388D(1) macrophages prelabeled with [(3)H]arachidonic acid (AA) respond to bacterial lipopolysaccharide (LPS) by mobilizing AA in a process that takes several hours and is mediated by the concerted actions of the group IV cytosolic phospholipase A(2) and the group V secretory phospholipase A(2) (sPLA(2)). Here we show that when the LPS-activated cells are prelabeled with [(3)H]oleic acid (OA), they also mobilize and release OA to the extracellular medium. The time and concentration dependence of the LPS effect on OA release fully resemble those of the AA release. Experiments in which both AA and OA release are measured simultaneously indicate that AA is released 3 times more efficiently than OA. Importantly, LPS-stimulated OA release is strongly inhibited by the selective sPLA(2) inhibitors 3-(3-acetamide-1-benzyl-2-ethylindolyl-5-oxy)propane sulfonic acid and carboxymethylcellulose-linked phosphatidylethanolamine. The addition of exogenous recombinant sPLA(2) to the cells also triggers OA release. These data implicate a functionally active sPLA(2) as being essential for the cells to release OA upon stimulation with LPS. OA release is also inhibited by methyl arachidonyl fluorophosphonate but not by bromoenol lactone, indicating that the group IV cytosolic phospholipase A(2) is also involved in the process. Together, these data reveal that OA release occurs during stimulation of the P388D(1) macrophages by LPS and that the regulatory features of the OA release are strikingly similar to those previously found for the AA release.  相似文献   

6.
P388D(1) cells exposed to bacterial lipopolysaccharide (LPS) mobilize arachidonic acid (AA) for prostaglandin synthesis in two temporally distinct pathways. The "immediate pathway" is triggered within minutes by receptor agonists such as platelet-activating factor (PAF) but only if the cells have previously been primed with LPS for 1 h. The "delayed pathway" occurs in response to LPS alone over the course of several hours. We have now investigated the subcellular localization of both the Group IV cytosolic phospholipase A(2) (cPLA(2)) and the Group V secreted PLA(2) (sPLA(2)) during these two temporally distinct routes of AA release. We have prepared cells overexpressing fusion proteins of sPLA(2)-GFP and cPLA(2)-RFP. In the resting cells, cPLA(2)-RFP was uniformly located throughout the cytoplasm, and short-term treatment with LPS did not induce translocation to perinuclear and/or Golgi membranes. However, such a translocation occurred almost immediately after the addition of PAF to the cells. Long-term exposure of the cells to LPS led to the translocation of cPLA(2)-RFP to intracellular membranes after 3 h, and correlates with a significant release of AA in a cPLA(2)-dependent manner. At the same time period that the delayed association of cPLA(2) with perinuclear membranes is detected, an intense fluorescence arising from the sPLA(2)-GFP was found around the nucleus in the sPLA(2)-GFP stably transfected cells. In parallel with these changes, significant AA release was detected from the sPLA(2)-GFP transfectants in a cPLA(2)-dependent manner, which may reflect cross-talk between sPLA(2) and cPLA(2). The subcellular localization of the Group VIA Ca(2+)-independent PLA(2) (iPLA(2)) was also investigated. Cells overexpressing iPLA(2)-GFP showed no fluorescence changes under any activation condition. However, the iPLA(2)-GFP-expressing cells showed relatively high basal AA release, confirming a role for iPLA(2) in basal deacylation reactions. These new data illustrate the subcellular localization changes that accompany the distinct roles that each of the three kinds of PLA(2) present in P388D(1) macrophages play in AA mobilization.  相似文献   

7.
Lipopolysaccharide (LPS) induces a delayed release (lag phase of 2-4 h) of arachidonic acid (AA) and prostaglandin (PG) D2 in rat liver macrophages. Group IV cytosolic phospholipase A2 (cPLA2) becomes phosphorylated within minutes after the addition of LPS. The phosphorylated form of cPLA2 shows an enhanced in vitro activity. The Ca2+ dependence of cPLA2 activity is not affected by phosphorylation of the enzyme. In addition, LPS induces an enhanced expression of cPLA2 mRNA (after 2-4 h) and an enhanced expression of cPLA2 protein (after 8 h). The cellular cPLA2 activity is enhanced about twofold 24 h after LPS treatment. Liver macrophages constitutively express mRNAs encoding Groups V and IIA secretory PLA2 (sPLA2). LPS has no effect on the levels of Groups V and IIA sPLA2 mRNA expression. Despite mRNA expression, Groups V and IIA sPLA2 protein and sPLA2 activity are not detectable in unstimulated or LPS-stimulated liver macrophages. Collectively, these and earlier [Mediators Inflammation 8 (1999) 295.] results suggest that in liver macrophages the LPS-induced delayed release of AA and prostanoids is mediated by phosphorylation and an enhanced expression of cPLA2, a de novo expression of cyclooxygenase (COX)-2, but not by the actions of Group V or Group IIA sPLA2.  相似文献   

8.
Previous studies have demonstrated that P388D(1) macrophages are able to mobilize arachidonic acid (AA) and synthesize prostaglandins in two temporally distinct phases. The first phase is triggered by platelet-activating factor within minutes, but needs the cells to be previously exposed to bacterial lipopolysaccharide (LPS) for periods up to 1 h. It is thus a primed immediate phase. The second, delayed phase occurs in response to LPS alone over long incubation periods spanning several hours. Strikingly, the effector enzymes involved in both of these phases are the same, namely the cytosolic group IV phospholipase A(2) (cPLA(2)), the secretory group V phospholipase A(2), and cyclooxygenase-2, although the regulatory mechanisms differ. Here we report that P388D(1) macrophages mobilize AA and produce prostaglandins in response to zymosan particles in a manner that is clearly different from the two described above. Zymosan triggers an immediate AA mobilization response from the macrophages that neither involves the group v phospholipase A(2) nor requires the cells to be primed by LPS. The group VI Ca(2+)-independent phospholipase A(2) is also not involved. Zymosan appears to signal exclusively through activation of the cPLA(2), which is coupled to the cyclooxygenase-2. These results define a secretory PLA(2)-independent pathway for AA mobilization in the P388D(1) macrophages, and demonstrate that, under certain experimental settings, stimulation of the cPLA(2) is sufficient to generate a prostaglandin biosynthetic response in the P388D(1) macrophages.  相似文献   

9.
The expression and regulation of the PGE receptors, EP(2) and EP(4), both of which are coupled to the stimulation of adenylate cyclase, were examined in peritoneal resident macrophages from C3H/HeN mice. mRNA expression of EP(4) but not EP(2) was found in nonstimulated cells, but the latter was induced by medium change alone, and this induction was augmented by LPS. mRNA expression of EP(4) was down-regulated by LPS but not by medium change. PGE(2) increased the cAMP content of both LPS-treated and nontreated cells. ONO-604, an EP(4) agonist, also increased cAMP content in nonstimulated cells and in cells treated with LPS for 3 h, but not for 6 h. Butaprost, an EP(2) agonist, was effective only in the cells treated with LPS for 6 h. The inhibitory effects of ONO-604 on TNF-alpha and IL-12 production were equipotent with PGE(2) at any time point, but the inhibitory effects of butaprost were only seen from 14 h after stimulation. PGE(2) or dibutyryl cAMP alone, but not butaprost, reduced EP(4) expression, and indomethacin reversed the LPS-induced down-regulation of EP(4), indicating that the down-regulation of EP(4) is mediated by LPS-induced PG synthesis and EP(4) activation. Indeed, when we used C3H/HeJ (LPS-hyporesponsive) macrophages, such reduction in EP(4) expression was found in the cells treated with PGE(2) alone, but not in LPS-treated cells. In contrast, up-regulation of EP(2) expression was again observed in LPS-treated C3H/HeJ macrophages. These results suggest that EP(4) is involved mainly in the inhibition of cytokine release, and that the gene expression of EP(2) and EP(4) is differentially regulated during macrophage activation.  相似文献   

10.
Group V secretory phospholipase A2 (sPLA2) rather than Group IIA sPLA2 is involved in short term, immediate arachidonic acid mobilization and prostaglandin E2 (PGE2) production in the macrophage-like cell line P388D1. When a new clone of these cells, P388D1/MAB, selected on the basis of high responsivity to lipopolysaccharide plus platelet-activating factor, was studied, delayed PGE2 production (6-24 h) in response to lipopolysaccharide alone occurred in parallel with the induction of Group V sPLA2 and cyclooxygenase-2 (COX-2). No changes in the level of cytosolic phospholipase A2 (cPLA2) or COX-1 were observed, and Group IIA sPLA2 was not detectable. Use of a potent and selective sPLA2 inhibitor, 3-(3-acetamide 1-benzyl-2-ethylindolyl-5-oxy)propanesulfonic acid (LY311727), and an antisense oligonucleotide specific for Group V sPLA2 revealed that delayed PGE2 was largely dependent on the induction of Group V sPLA2. Also, COX-2, not COX-1, was found to mediate delayed PGE2 production because the response was completely blocked by the specific COX-2 inhibitor NS-398. Delayed PGE2 production and Group V sPLA2 expression were also found to be blunted by the inhibitor methylarachidonyl fluorophosphonate. Because inhibition of Ca2+-independent PLA2 by an antisense technique did not have any effect on the arachidonic acid release, the data using methylarachidonyl fluorophosphonate suggest a key role for the cPLA2 in the response as well. Collectively, the results suggest a model whereby cPLA2 activation regulates Group V sPLA2 expression, which in turn is responsible for delayed PGE2 production via COX-2.  相似文献   

11.
In order to ascertain the role of phospholipase A2 (PLA2) in the release of arachidonic acid for eicosanoid biosynthesis, we have characterized a Ca2+-dependent PLA2 from P388D1 cells, evaluated inhibitors of its activity, and correlated the effects of these inhibitors on prostaglandin (PG) E2 production in the intact cell. The Ca2+-dependent PLA2 has little preference for the polar head group or sn-2 fatty acid of phospholipids, and we have now found that it will hydrolyze 1-alkyl,2-acyl phospholipids, but it does not show a preference for this substrate over other phospholipids. Inhibitor studies with the Ca2+-dependent PLA2 have shown that arachidonic acid is an effective inhibitor. The analogs of natural fatty acids, eicosatetraynoic acid and octadecyleicosaynoic acid, were ineffective as inhibitors of the P388D1 PLA2. However, 7,7-dimethyl-5,8-eicosadienoic acid was as effective an inhibitor (IC50 = 16 microM) as arachidonic acid. Manoalide and its analog, manoalogue, were found to be good inhibitors of the P388D1 PLA2 (IC50 = 16 and 26 microM, respectively). The irreversible inhibitor of the extracellular PLA2, p-bromophenacyl bromide, was a very poor inhibitor of the P388D1 PLA2, apparent IC50 = 500-600 microM. Quinacrine was also ineffective as an inhibitor as was the cyclooxygenase inhibitor indomethacin. On the cellular level, the P388D1 cells respond to various stimuli to produce PGD2 and PGE2 as the major cyclooxygenase products with minor production of PGI2 and thromboxane A2. Similar arachidonic acid metabolite profiles were seen for calcium ionophore A23187, melittin, and platelet-activating factor. Manoalide, manoalogue, and 7,7-dimethyl-5,8-eicosadienoic acid, effective inhibitors of the isolated PLA2, inhibited PGE2 production in intact P388D1 cells 40-85% in the concentration range studied. In contrast, p-bromophenacyl bromide, which is ineffective as an inhibitor of the P388D1 PLA2, did not significantly effect PGE2 production in the concentration ranges used. These results demonstrate that there may be important differences between the intracellular P388D1 PLA2 and the more commonly studied extracellular forms of PLA2. These differences are also observed in the intact cell studies and emphasize the need for the evaluation of inhibitors both in vitro and in vivo using the isolated enzyme and intact cell. This is the first example of studies aimed at correlating the inhibition of a purified intracellular PLA2 with inhibition of prostaglandin production in the intact cell from which it is derived.  相似文献   

12.
Ceramide has been suggested to function as a mediator of exocytosis in response to the addition of a calcium ionophore from PC12 cells. Here, we show that although cell-permeable C(6)-ceramide or a calcium ionophore alone did not increase either the degranulation of serotonin or the release of arachidonic acid (AA) from RBL-2H3 cells, their combined effect significantly stimulated these processes in a time- and dose-dependent manner. This effect was inhibited by the presence of an exogenous calcium chelator and significantly suppressed by the CERK inhibitor (K1) and phospholipase A(2) (PLA(2)) inhibitors. Moreover, cytosolic PLA(2) GIVA (cPLA(2) GIVA) siRNA-transfected RBL-2H3 cells showed a lower level of serotonin release than scramble siRNA-transfected cells. Little is known about the regulation of degranulation proximal to the activation of cytosolic phospholipase A(2) GIVA, the initial rate-limiting step in RBL-2H3 cells. In this study, we suggest that CERK, ceramide-1-phosphate, and PLA(2) are involved in degranulation in a calcium-dependent manner. Inhibition of p44/p42 mitogen-activated protein kinase partially decreased the AA release, but did not affect degranulation. Furthermore, treatment of the cells with AA (ω-6, C20:4), not linoleic acid (ω-6, C18:2) or α-linolenic acid (ω-6, C18:3), induced degranulation. Taken together, these results suggest that ceramide is involved in mast cell degranulation via the calcium-mediated activation of PLA(2).  相似文献   

13.
Group IVA phospholipase A2 (GIVA PLA2) catalyzes the release of arachidonic acid (AA) from the sn-2 position of glycerophospholipids. AA is then further metabolized into terminal signaling molecules including numerous prostaglandins. We have now demonstrated the involvement of phosphatidic acid phosphohydrolase 1 (PAP-1) and protein kinase C (PKC) in the Toll-like receptor-4 (TLR-4) activation of GIVA PLA2. We also studied the effect of PAP-1 and PKC on Ca+ 2 induced and synergy enhanced GIVA PLA2 activation. We observed that the AA release induced by exposure of RAW 264.7 macrophages to the TLR-4 specific agonist Kdo2-Lipid A is blocked by the PAP-1 inhibitors bromoenol lactone (BEL) and propranolol as well as the PKC inhibitor Ro 31-8220; however these inhibitors did not reduce AA release stimulated by Ca+ 2 influx induced by the P2X7 purinergic receptor agonist ATP. Additionally, stimulation of cells with diacylglycerol (DAG), the product of PAP-1 mediated hydrolysis, initiated AA release from unstimulated cells as well as restored normal AA release from cells treated with PAP-1 inhibitors. Finally, neither PAP-1 nor PKC inhibition reduced GIVA PLA2 synergistic activation by stimulation with Kdo2-Lipid A and ATP.  相似文献   

14.
Cytosolic phospholipase A2 (cPLA2) plays a pivotal role in mediating agonist-induced arachidonic acid (AA) release for prostaglandins (PG) synthesis induced by bacterial lipopolysaccharide (LPS) and cytokines. However, the intracellular signaling pathways mediating LPS-induced cPLA2 expression and PGE2 synthesis in canine tracheal smooth muscle cells (TSMCs) remains unknown. LPS-induced expression of cPLA2 and release of PGE2 was attenuated by inhibitors of tyrosine kinase (genistein), phosphatidylcholine-phospholipase C (D609), phosphatidylinositol-phospholipase C (U73122), PKC (GF109203X and staurosporine), removal of Ca2+ by BAPTA/AM plus EDTA, MEK1/2 (PD98059), p38 (SB202190), JNK (SP600125), and phosphatidylinositol 3-kinase (PI3-K; LY294002 and wortmannin). The involvement of MPAKs in LPS-induced responses was further confirmed by transfection of TSMCs with dominant negative mutants of ERK2 and p38. LPS-induced cPLA2 expression and PGE2 synthesis was inhibited by a selective NF-kappaB inhibitor (helenalin) and transfection with dominant negative mutants of NF-kappaB inducing kinase (NIK), IkappaB kinase (IKK)-alpha, and IKK-beta, consistent with that LPS-stimulated both IkappaB-alpha degradation and NF-kappaB translocation into nucleus in these cells. LPS-stimulated cPLA2 phosphorylation was inhibited by PD98059, GF109203X, and staurosporine, indicating the regulation by p42/p44 MAPK and PKC. Moreover, LPS-induced up-regulation of cPLA2 and COX-2 linked to PGE2 synthesis was inhibited by AACOCF3 (a selective cPLA2 inhibitor), implying the involvement of cPLA2 in these responses. These findings suggest that phosphorylation and expression of cPLA2 correlates with the release of PGE2 from LPS-challenged TSMCs, at least in part, mediated through MAPKs and NF-kappaB signaling pathways. LPS-mediated responses were modulated by PLC, Ca2+, PKC, tyrosine kinase, and PI3-K in TSMCs.  相似文献   

15.
16.
The small intestinal epithelium is a highly dynamic system continuously renewed by a process involving cell proliferation and differentiation. The intestinal epithelium constitutes a permeability barrier regulating the vectorial transport of ions, water, and solutes. Morphological changes during cell differentiation, as well as changes in the activity of brush-border enzymes and the expression of transport proteins, are well established. However, little is known about the arachidonic acid (AA) cascade underlying epithelial cell differentiation or its role in the development of epithelial barrier function. The main purpose of this study was to examine the activity of the high-molecular-weight phospholipases A(2) (PLA(2)) and cyclooxygenase (COX) pathway during differentiation, with particular emphasis on paracellular permeability. PLA(2) activity, AA release, COX-2 expression, prostaglandin E(2) (PGE(2)) production, and paracellular permeability were studied in preconfluent, confluent, and differentiated Caco-2 cell cultures. Our results show that Caco-2 differentiation induces a decrease in both calcium-independent PLA(2) activity and COX-2 expression and, consequently, a decrease in AA release and PGE(2) synthesis in parallel with a reduction in paracellular permeability. Moreover, the addition of PGE(2) to differentiated cells, at concentrations similar to those detected in nondifferentiated cultures, induces the disruption of epithelial barrier function. These results suggest that AA release by calcium-independent PLA(2), COX-2 expression, and subsequent PGE(2) release are important for the maintenance of paracellular permeability in differentiated Caco-2 cells.  相似文献   

17.
We determined the roles of reactive oxygen species (ROS) in the expression of cyclooxygenase-2 (COX-2) and the production of prostaglandin E2 (PGE2) in lipopolysaccharide (LPS)-activated microglia. LPS treatment increased intracellular ROS in rat microglia dose-dependently. Pre-treatment with superoxide dismutase (SOD)/catalase, or SOD/catalase mimetics that can scavenge intracellular ROS, significantly attenuated LPS-induced release in PGE2. Diphenylene iodonium (DPI), a non-specific NADPH oxidase inhibitor, decreased LPS-induced PGE2 production. In addition, microglia from NADPH oxidase-deficient mice produced less PGE2 than those from wild-type mice following LPS treatment. Furthermore, LPS-stimulated expression of COX-2 (determined by RT-PCR analysis of COX-2 mRNA and western blot for its protein) was significantly reduced by pre-treatment with SOD/catalase or SOD/catalase mimetics. SOD/catalase mimetics were more potent than SOD/catalase in reducing COX-2 expression and PGE2 production. As a comparison, scavenging ROS had no effect on LPS-induced nitric oxide production in microglia. These results suggest that ROS play a regulatory role in the expression of COX-2 and the subsequent production of PGE2 during the activation process of microglia. Thus, inhibiting NADPH oxidase activity and subsequent ROS generation in microglia can reduce COX-2 expression and PGE2 production. These findings suggest a potential therapeutic intervention strategy for the treatment of inflammation-mediated neurodegenerative diseases.  相似文献   

18.
Conditioned medium from P388 D1 cell line containing interleukin 1 (IL-1) and granulocyte macrophage colony stimulating factor (GM-CSF) can stimulate prostaglandin E2 (PGE2) production by murine bone marrow cells. In this work, we show that although GM-CSF (either purified from P388 D1 CM or murine recombinant GM-CSF) does not significantly alter bone marrow cell PGE2 production, its presence in P388 D1 CM is however necessary to induce this effect since the presence of anti GM-CSF antiserum completely abrogated the increase in PGE2 production in response to P388 D1 CM. In addition IL-1 tested alone does not not modify PGE2 release by bone marrow cells. However, the simultaneous addition of IL-1 and GM-CSF markedly increases PGE2 production. Thus, the ability of P388 D1 CM to stimulate PGE2 synthesis by bone marrow cells appears to result from a synergistic action between GM-CSF and IL-1.  相似文献   

19.
Phospholipases A2 (PLA2) and cyclooxygenases (COX) are important enzymes responsible for production of potent lipid mediators, including prostaglandins (PG) and thromboxane A2. We investigated coupling between PLA2 and COX isoforms by using transient transfection in COS-1 cells. Untransfected cells, incubated with or without phorbol ester + the Ca2+ ionophore ionomycin, generated trivial amounts of PGE2. In cells co-transfected with cytosolic PLA2 (cPLA2) and COX-1 or COX-2, phorbol ester + ionomycin markedly stimulated PGE2 production. There was no preferential coupling of cPLA2 to either of the COX isoforms. In contrast, group IIA secretory PLA2 (sPLA2) co-transfected with COX-1 or COX-2 did not lead to an increase in PGE2 production, despite high levels of sPLA2 enzymatic activity. Transfection of cPLA2 did not affect basal free arachidonic acid (AA) levels. Phorbol ester + ionomycin stimulated release of AA in cPLA2-transfected COS-1 cells, but not in untransfected cells, whereas sPLA2 transfection (without stimulation) led to high basal free AA. Thus, AA released by cPLA2 is accessible to both COX isoforms for metabolism to PG, whereas AA released by sPLA2 is not metabolized by COX.  相似文献   

20.
Lipopolysaccharide (LPS) was found to induce inflammatory responses in the airways and exerted as a potent stimulus for PG synthesis. This study was to determine the mechanisms of LPS-enhanced cyclooxygenase (COX)-2 expression associated with PGE(2) synthesis in tracheal smooth muscle cells (TSMCs). LPS markedly increased the expression of COX-2 and release of PGE(2) in a time- and concentration-dependent manner, whereas COX-1 remained unaltered. Both the expression of COX-2 and the generation of PGE(2) in response to LPS were attenuated by a tyrosine kinase inhibitor genistein, a phosphatidylcholine-phospholipase C inhibitor D609, a phosphatidylinositol-phospholipase C inhibitor U73122, protein kinase C inhibitors, GF109203X and staurosporine, removal of Ca(2+) by addition of BAPTA/AM plus EGTA, and phosphatidylinositol 3-kinase (PI3-K) inhibitors, LY294002 and wortmannin. Furthermore, LPS-induced NF-kappaB activation correlated with the degradation of IkappaB-alpha, COX-2 expression, and PGE(2) synthesis, was inhibited by transfection with dominant negative mutants of NIK and IKK-alpha, but not by IKK-beta. LPS-induced COX-2 expression and PGE(2) synthesis were completely inhibited by PD98059 (an inhibitor of MEK1/2) and SB203580 (an inhibitor of p38 MAPK inhibitor), but these two inhibitors had no effect on LPS-induced NF-kappaB activation, indicating that NF-kappaB is activated by LPS independently of activation of p42/p44 MAPK and p38 MAPK pathways in TSMCs. Taken together, these findings suggest that the increased expression of COX-2 correlates with the release of PGE(2) from LPS-challenged TSMCs, at least in part, independently mediated through MAPKs and NF-kappaB signalling pathways. LPS-mediated responses were modulated by PLC, Ca(2+), PKC, tyrosine kinase, and PI3-K in these 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