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 46 毫秒
1.
Although it has been proposed that arachidonate release by several secretory phospholipase A2 (sPLA2) isozymes is modulated by cytosolic PLA2 (cPLA2), the cellular component(s) that intermediates between these two signaling PLA2s remains unknown. Here we provide evidence that 12- or 15-lipoxygenase (12/15-LOX), which lies downstream of cPLA2, plays a pivotal role in cytokine-induced gene expression and function of sPLA2-IIA. The sPLA2-IIA expression and associated PGE2 generation induced by cytokines in rat fibroblastic 3Y1 cells were markedly attenuated by antioxidants that possess 12/15-LOX inhibitory activity. 3Y1 cells expressed 12/15-LOX endogenously, and forcible overexpression of 12/15-LOX in these cells greatly enhanced cytokine-induced expression of sPLA2-IIA, with a concomitant increase in delayed PG generation. Moreover, studies using 293 cells stably transfected with sPLA2-IIA revealed that stimulus-dependent hydrolysis of membrane phospholipids by sPLA2-IIA was enhanced by overexpression of 12/15-LOX. These results indicate that the product(s) generated by the cPLA2-12/15-LOX pathway following cell activation may play two roles: enhancement of sPLA2-IIA gene expression and membrane sensitization that leads to accelerated sPLA2-IIA-mediated hydrolysis.  相似文献   

2.
We previously described that recombinant interleukin-1beta (IL-1beta) induced the significant release of substance P (SP) via a cyclooxygenase (COX) pathway in primary cultured rat dorsal root ganglion (DRG) cells. In the present study, we examined the involvement of two types of phospholipase A2 (PLA2) enzymes, which lie upstream of COX in the prostanoid-generating pathway, in the IL-1beta-induced release of SP from DRG cells. The expression of type IIA secretory PLA2 (sPLA2 -IIA) mRNA was undetectable by ribonuclease protection assay in non-treated DRG cells, while in DRG cells incubated with 1 ng/mL of IL-1beta, the expression was induced in a time-dependent manner. On the other hand, type IV cytosolic PLA2 (cPLA2 ) mRNA was constitutively expressed in the non-treated DRG cells, and treatment with 1 ng/mL of IL-1beta for 3 h significantly increased the levels of cPLA2 mRNA. The IL-1beta-induced SP release was significantly inhibited by the sPLA2 inhibitor, thioetheramide phosphorylcholine (TEA-PC), and the cPLA2 inhibitor, arachidonyl trifluoromethyl ketone (AACOCF3 ). Furthermore AACOCF3 suppressed the induction of sPLA2 -IIA mRNA expression induced by IL-1beta. These observations suggested that two types of PLA2, sPLA2 -IIA and cPLA2, were involved in the IL-1beta-induced release of SP from DRG cells, and that the functional cross-talk between the two enzymes might help to control their activity in the prostanoid-generating system in DRG cells. These events might be key steps in the inflammation-induced hyperactivity in primary afferent neurons of spinal cord.  相似文献   

3.
T Kambe  M Murakami  I Kudo 《FEBS letters》1999,453(1-2):81-84
By analyzing human embryonic kidney 293 cell transfectants stably overexpressing various types of phospholipase A2 (PLA2), we have shown that polyunsaturated fatty acids (PUFAs) preferentially activate type IIA secretory PLA2 (sPLA2-IIA)-mediated arachidonic acid (AA) release from interleukin-1 (IL-1)-stimulated cells. When 293 cells prelabeled with 13H]AA were incubated with exogenous PUFAs in the presence of IL-1 and serum, there was a significant increase in [3H]AA release (in the order AA > linoleic acid > oleic acid), which was augmented markedly by sPLA2-IIA and modestly by type IV cytosolic PLA2 (cPLA2), but only minimally by type VI Ca2(+)-independent PLA2, overexpression. Transfection of cPLA2 into sPLA2-IIA-expressing cells produced a synergistic increase in IL-1-dependent [3H]AA release and subsequent prostaglandin production. Our results support the proposal that prior production of AA by cPLA2 in cytokine-stimulated cells destabilizes the cellular membranes, thereby rendering them more susceptible to subsequent hydrolysis by sPLA2-IIA.  相似文献   

4.
Phosphatidylcholine (PtdCho) is a major membrane phospholipid, and its loss is sufficient in itself to induce cell death. PtdCho homeostasis is regulated by the balance between hydrolysis and synthesis. PtdCho is hydrolyzed by phospholipase A2 (PLA2), PtdChospecific phospholipase C (PtdCho-PLC), and phospholipase D (PLD). PtdCho synthesis is rate-limited by CTP:phosphocholine cytidylyltransferase (CCT), which makes CDP-choline. The final step of PtdCho synthesis is catalyzed by CDP-choline:1,2-diacylglycerol cholinephosphotransferase. PtdCho synthesis in the brain is predominantly through the CDP-choline pathway. Transient middle cerebral artery occlusion (tMCAO) significantly increased PLA2 activity, secretory PLA2 (sPLA2)-IIA mRNA and protein levels, PtdCho-PLC activity, and PLD2 protein expression following reperfusion. CDP-choline treatment significantly attenuated PLA2 activity, sPLA2-IIA mRNA and protein levels, and PtdCho-PLC activity, but did not affect PLD2 protein expression. tMCAO also resulted in loss of CCT activity and CCTalpha protein, which were partially restored by CDP-choline. No changes were observed in cytosolic PLA2 or calcium-independent PLA2 tMCAO. protein levels after Up-regulation of PLA2, PtdCho-PLC, and PLD and regulation of CCT collectively down-resulted in loss of PtdCho, which was significantly restored by CDP-choline treatment. CDP-choline treatment significantly attenuated the infarction volume by 55 +/- 5% after 1 h of tMCAO and 1 day of reperfusion. Taken together, these results suggest that CDP-choline significantly restores Ptd-Cho levels by differentially affecting sPLA2-IIA, PtdCho-PLC, and CCTalpha after transient focal cerebral ischemia. A hypothetical scheme is proposed integrating results from this study and from other reports in the literature.  相似文献   

5.
Accumulating evidence has suggested that cytosolic phospholipase A(2) (cPLA(2)) and several secretory PLA(2) (sPLA(2)) isozymes are signaling PLA(2)s that are functionally coupled with downstream cyclooxygenase (COX) isozymes for prostaglandin (PG) biosynthesis. Arachidonic acid (AA) released by cPLA(2) and sPLA(2)s is supplied to both COX-1 and COX-2 in the immediate, and predominantly to COX-2 in the delayed, PG-biosynthetic responses. Vimentin, an intermediate filament component, acts as a functional perinuclear adapter for cPLA(2), in which the C2 domain of cPLA(2) associates with the head domain of vimentin in a Ca(2+)-sensitive manner. The heparin-binding signaling sPLA(2)-IIA, IID and V bind the glycosylphosphatidylinositol-anchored heparan sulfate proteoglycan glypican, which plays a role in sorting of these isozymes into caveolae and perinuclear compartments. Phospholipid scramblase, which facilitates transbilayer movement of anionic phospholipids, renders the cellular membranes more susceptible to signaling sPLA(2)s. There is functional cooperation between cPLA(2) and signaling sPLA(2)s in that prior activation of cPLA(2) is required for the signaling sPLA(2)s to act properly. cPLA(2)-derived AA is oxidized by 12/15-lipoxygenase, the products of which not only augment the induction of sPLA(2) expression, but also cause membrane perturbation, leading to increased cellular susceptibility to the signaling sPLA(2)s. sPLA(2)-X, a heparin-non-binding sPLA(2) isozyme, is capable of releasing AA from intact cells in the absence of cofactors. This property is attributed to its ability to avidly hydrolyze zwitterionic phosphatidylcholine, a major phospholipid in the outer plasma membrane. sPLA(2)-V can also utilize this route in several cell types. Taken together, the AA-releasing function of sPLA(2)s depends on the presence of regulatory cofactors and interfacial binding to membrane phospholipids, which differ according to cell type, stimuli, secretory processes, and subcellular distributions.  相似文献   

6.
Group IIA secretory phospholipase A(2) (sPLA(2)-IIA) is a prototypic sPLA(2) enzyme that may play roles in modification of eicosanoid biosynthesis as well as antibacterial defense. In several cell types, inducible expression of sPLA(2) by pro-inflammatory stimuli is attenuated by group IVA cytosolic PLA(2) (cPLA(2)alpha) inhibitors such as arachidonyl trifluoromethyl ketone, leading to the proposal that prior activation of cPLA(2)alpha is required for de novo induction of sPLA(2). However, because of the broad specificity of several cPLA(2)alpha inhibitors used so far, a more comprehensive approach is needed to evaluate the relevance of this ambiguous pathway. Here, we provide evidence that the induction of sPLA(2)-IIA by pro-inflammatory stimuli requires group VIB calcium-independent PLA(2) (iPLA(2)gamma), rather than cPLA(2)alpha, in rat fibroblastic 3Y1 cells. Results with small interfering RNA unexpectedly showed that the cytokine induction of sPLA(2)-IIA in cPLA(2)alpha knockdown cells, in which cPLA(2)alpha protein was undetectable, was similar to that in replicate control cells. By contrast, knockdown of iPLA(2)gamma, another arachidonyl trifluoromethyl ketone-sensitive intracellular PLA(2), markedly reduced the cytokine-induced expression of sPLA(2)-IIA. Supporting this finding, the R-enantiomer of bromoenol lactone, an iPLA(2)gamma inhibitor, suppressed the cytokine-induced sPLA(2)-IIA expression, whereas (S)-bromoenol lactone, an iPLA(2)beta inhibitor, failed to do so. Moreover, lipopolysaccharide-stimulated sPLA(2)-IIA expression was also abolished by knockdown of iPLA(2)gamma. These findings open new insight into a novel regulatory role of iPLA(2)gamma in stimulus-coupled sPLA(2)-IIA expression.  相似文献   

7.
Human group IIA-secreted phospholipase A(2) (sPLA(2)-IIA) is an important regulator of cytokine-mediated inflammatory responses in both in vitro and in vivo models of rheumatoid arthritis (RA). However, treatment of RA patients with sPLA(2)-IIA inhibitors shows only transient benefit. Using an activity-impaired sPLA(2)-IIA mutant protein (H48Q), we show that up-regulation of TNF-dependent PGE(2) production and cyclooxygenase-2 (COX-2) induction by exogenous sPLA(2)-IIA in RA fibroblast-like synoviocytes (FLSs) is independent of its enzyme function. Selective cytosolic phospholipase A(2)-α (cPLA(2)-α) inhibitors abrogate TNF/sPLA(2)-IIA-mediated PGE(2) production without affecting COX-2 levels, indicating arachidonic acid (AA) flux to COX-2 occurs exclusively through TNF-mediated activation of cPLA(2)-α. Nonetheless, exogenous sPLA(2)-IIA, but not H48Q, stimulates both AA mobilization from FLSs and microparticle-derived AA release that is not used for COX-2-dependent PGE(2) production. sPLA(2)-IIA-mediated AA production is inhibited by pharmacological blockade of sPLA(2)-IIA but not cPLA(2)-α. Exogenous H48Q alone, like sPLA(2)-IIA, increases COX-2 protein levels without inducing PGE(2) production. Unlike TNF, sPLA(2)-IIA alone does not rapidly mobilize NF-κB or activate phosphorylation of p38 MAPK, two key regulators of COX-2 protein expression, but does activate the ERK1/2 pathway. Thus, sPLA(2)-IIA regulates AA flux through the cPLA(2)-α/COX-2 pathway in RA FLSs by up-regulating steady state levels of these biosynthetic enzymes through an indirect mechanism, rather than direct provision of substrate to the pathway. Inhibitors that have been optimized for their potency in enzyme activity inhibition alone may not adequately block the activity-independent function of sPLA(2)-IIA.  相似文献   

8.
Activation of brain mitochondrial phospholipase(s) A(2) (PLA(2)) might contribute to cell damage and be involved in neurodegeneration. Despite the potential importance of the phenomenon, the number, identities, and properties of these enzymes are still unknown. Here, we demonstrate that isolated mitochondria from rat brain cortex, incubated in the absence of respiratory substrates, release a Ca(2+)-dependent PLA(2) having biochemical properties characteristic to secreted PLA(2) (sPLA(2)) and immunoreacting with the antibody raised against recombinant type IIA sPLA(2) (sPLA(2)-IIA). Under identical conditions, no release of fumarase in the extramitochondrial medium was observed. The release of sPLA(2) from mitochondria decreases when mitochondria are incubated in the presence of respiratory substrates such as ADP, malate, and pyruvate, which causes an increase of transmembrane potential determined by cytofluorimetric analysis using DiOC(6)(3) as a probe. The treatment of mitochondria with the uncoupler carbonyl cyanide 3-chlorophenylhydrazone slightly enhances sPLA(2) release. The increase of sPLA(2) specific activity after removal of mitochondrial outer membrane indicates that the enzyme is associated with mitoplasts. The mitochondrial localization of the enzyme has been confirmed by electron microscopy in U-251 astrocytoma cells and by confocal laser microscopy in the same cells and in PC-12 cells, where the structurally similar isoform type V-sPLA(2) has mainly nuclear localization. In addition to sPLA(2), mitochondria contain another phospholipase A(2) that is Ca(2+)-independent and sensitive to bromoenol lactone, associated with the outer mitochondrial membrane. We hypothesize that, under reduced respiratory rate, brain mitochondria release sPLA(2)-IIA that might contribute to cell damage.  相似文献   

9.
In macrophages and other major immunoinflammatory cells, two phospholipase A(2) (PLA(2)) enzymes act in concert to mobilize arachidonic acid (AA) for immediate PG synthesis, namely group IV cytosolic phospholipase A(2) (cPLA(2)) and a secreted phospholipase A(2) (sPLA(2)). In this study, the molecular mechanism underlying cross-talk between the two PLA(2)s during paracrine signaling has been investigated. U937 macrophage-like cells respond to Con A by releasing AA in a cPLA(2)-dependent manner, and addition of exogenous group V sPLA(2) to the activated cells increases the release. This sPLA(2) effect is abolished if the cells are pretreated with cPLA(2) inhibitors, but is restored by adding exogenous free AA. Inhibitors of cyclooxygenase and 5-lipoxygenase have no effect on the response to sPLA(2). In contrast, ebselen strongly blocks it. Reconstitution experiments conducted in pyrrophenone-treated cells to abolish cPLA(2) activity reveal that 12- and 15-hydroperoxyeicosatetraenoic acid (HPETE) are able to restore the sPLA(2) response to levels found in cells displaying normal cPLA(2) activity. Moreover, 12- and 15-HPETE are able to enhance sPLA(2) activity in vitro, using a natural membrane assay. Neither of these effects is mimicked by 12- or 15-hydroxyeicosatetraenoic acid, indicating that the hydroperoxy group of HPETE is responsible for its biological activity. Collectively, these results establish a role for 12/15-HPETE as an endogenous activator of sPLA(2)-mediated phospholipolysis during paracrine stimulation of macrophages and identify the mechanism that connects sPLA(2) with cPLA(2) for a full AA mobilization response.  相似文献   

10.
Proteinase-activated receptor-1 (PAR1), upon activation, exerts prostanoid-dependent gastroprotection, and increases prostaglandin E(2) (PGE(2)) release through cyclooxygenase-2 (COX-2) upregulation in rat gastric mucosal epithelial RGM1 cells. However, there is a big time lag between the PAR1-triggered PGE(2) release and COX-2 upregulation in RGM1 cells; that is, the former event takes 18 h to occur, while the latter rapidly develops and reaches a plateau in 6 h. The present study thus aimed at clarifying mechanisms for the delay of PGE(2) release after PAR1 activation in RGM1 cells. Although a PAR1-activating peptide, TFLLR-NH(2), alone caused PGE(2) release at 18 h, but not 6 h, TFLLR-NH(2) in combination with arachidonic acid dramatically enhanced PGE(2) release even for 1-6 h. TFLLR-NH(2) plus linoleic acid caused a similar rapid response. CP-24879, a Δ(5)/Δ(6)-desaturase inhibitor, abolished the PGE(2) release induced by TFLLR-NH(2) plus linoleic acid, but not by TFLLR-NH(2) alone. The TFLLR-NH(2)-induced PGE(2) release was not affected by inhibitors of cytosolic phospholipase A(2) (cPLA(2)), Ca(2+)-independent PLA(2) (cPLA(2)) or secretory PLA(2) (sPLA(2)), but was abolished by their mixture or a pan-PLA(2) inhibitor. Among PLA(2) isozymes, mRNA of group IIA sPLA(2) (sPLA(2)-IIA) was upregulated following PAR1 stimulation for 6-18 h, whereas protein levels of PGE synthases were unchanged. These data suggest that the delay of PGE(2) release after COX-2 upregulation triggered by PAR1 is due to the poor supply of free arachidonic acid at the early stage in RGM1 cells, and that plural isozymes of PLA(2) including sPLA(2)-IIA may complementarily contribute to the liberation of free arachidonic acid.  相似文献   

11.
12.
We herein demonstrate that mast cells express all known members of the group II subfamily of secretory phospholipase A2 (sPLA2) isozymes, and those having heparin affinity markedly enhance the exocytotic response. Rat mastocytoma RBL-2H3 cells transfected with heparin-binding (sPLA2-IIA, -V, and -IID), but not heparin-nonbinding (sPLA2-IIC), enzymes released more granule-associated markers (beta-hexosaminidase and histamine) than mock- or cytosolic PLA2alpha (cPLA2alpha)-transfected cells after stimulation with IgE and Ag. Site-directed mutagenesis of sPLA2-IIA and -V revealed that both the catalytic and heparin-binding domains are essential for this function. Confocal laser and electron microscopic analyses revealed that sPLA2-IIA, which was stored in secretory granules in unstimulated cells, accumulated on the membranous sites where fusion between the plasma membrane and granule membranes occurred in activated cells. These results suggest that the heparin-binding sPLA2s bind to the perigranular membranes through their heparin-binding domain, and lysophospholipids produced in situ by their enzymatic action may facilitate the ongoing membrane fusion. In contrast to the redundant role of sPLA2-IIA, -IID, and -V in the regulation of degranulation, only sPLA2-V had the ability to markedly augment IgE/Ag-stimulated immediate PGD2 production, which reached a level comparable to that elicited by cPLA2alpha. The latter observation reveals an unexplored functional segregation among the three related isozymes expressed in the same cell population.  相似文献   

13.
The purpose of this study was to define the role of secretory phospholipase A2 (sPLA2), calcium-independent PLA2, and cytosolic PLA2 (cPLA2) in arachidonic acid (AA) release from fMLP-stimulated human neutrophils. While fMLP induced the release of extracellular sPLA2 activity and AA, 70% of sPLA2 activity remained associated with the cell. Treatment with the cell-impermeable sPLA2 inhibitors DTT or LY311-727, or the anti-sPLA2 Ab 3F10 all inactivated extracellular sPLA2 activity, but had minimal effect on neutrophil AA mass release. In contrast, coincubation of streptolysin-O toxin-permeabilized neutrophils with DTT, LY311-727, or 3F10 all decreased [3H8]AA release from [3H8]AA-labeled, fMLP-stimulated cells. Exposure to fMLP resulted in a decrease in the electrophoretic mobility of cPLA2, a finding consistent with cPLA2 phosphorylation, and stimulated the translocation of cPLA2 from cytosolic to microsomal and nuclear compartments. The role of cPLA2 was further evaluated with the cPLA2 inhibitor methyl arachidonyl fluorophosphonate, which attenuated cPLA2 activity in vitro and decreased fMLP-stimulated AA mass release by intact neutrophils, but had no effect on neutrophil sPLA2 activity. Inhibition of calcium-independent PLA2 with haloenol lactone suicide substrate had no effect on neutrophil cPLA2 activity or AA mass release. These results indicate a role for cPLA2 and an intracellular or cell-associated sPLA2 in the release of AA from fMLP-stimulated human neutrophils.  相似文献   

14.
Oxidation and lipolytic remodeling of LDL are believed to stimulate LDL entrapment in the arterial wall, expanding the inflammatory response and promoting atherosclerosis. However, the cellular responses and molecular mechanisms underlying the atherogenic effects of lipolytically modified LDL are incompletely understood. Human THP-1 monocytes were prelabeled with [(3)H]arachidonic acid (AA) before incubation with LDL or LDL lipolytically modified by secretory PLA(2) (sPLA(2)) or bacterial sphingomyelinase (SMase). LDL elicited rapid and dose-dependent extracellular release of AA in monocytes. Interestingly, LDL modified by sPLA(2) or SMase displayed a marked increase in AA mobilization relative to native LDL, and this increase correlated with enhanced activity of cytosolic PLA(2) (cPLA(2)) assayed in vitro as well as increased monocyte tumor necrosis factor-alpha secretion. The AA liberation was attenuated by inhibitors toward cPLA(2) and sPLA(2), indicating that both PLA(2) enzymes participate in LDL-induced AA release. In conclusion, these results demonstrate that LDL lipolytically modified by sPLA(2) or SMase potentiates cellular AA release and cPLA(2) activation in human monocytes. From our results, we suggest novel atherogenic properties for LDL modified by sPLA(2) and SMase in AA release and signaling, which could contribute to the inflammatory gene expression observed in atherosclerosis.  相似文献   

15.
We analyzed the ability of a diverse set of mammalian secreted phospholipase A(2) (sPLA(2)) to release arachidonate for lipid mediator generation in two transfected cell lines. In human embryonic kidney 293 cells, the heparin-binding enzymes sPLA(2)-IIA, -IID, and -V promote stimulus-dependent arachidonic acid release and prostaglandin E(2) production in a manner dependent on the heparan sulfate proteoglycan glypican. In contrast, sPLA(2)-IB, -IIC, and -IIE, which bind weakly or not at all to heparanoids, fail to elicit arachidonate release, and addition of a heparin binding site to sPLA(2)-IIC allows it to release arachidonate. Heparin nonbinding sPLA(2)-X liberates arachidonic acid most likely from the phosphatidylcholine-rich outer plasma membrane in a glypican-independent manner. In rat mastocytoma RBL-2H3 cells that lack glypican, sPLA(2)-V and -X, which are unique among sPLA(2)s in being able to hydrolyze phosphatidylcholine-rich membranes, act most likely on the extracellular face of the plasma membrane to markedly augment IgE-dependent immediate production of leukotriene C(4) and platelet-activating factor. sPLA(2)-IB, -IIA, -IIC, -IID, and -IIE exert minimal effects in RBL-2H3 cells. These results are also supported by studies with sPLA(2) mutants and immunocytostaining and reveal that sPLA(2)-dependent lipid mediator generation occur by distinct (heparanoid-dependent and -independent) mechanisms in HEK293 and RBL-2H3 cells.  相似文献   

16.
Recently, a novel peptide (Trp-Lys-Tyr-Met-Val-D-Met, WKYMVm) has been shown to induce superoxide generation in human monocytes. The peptide stimulated phospholipase A2 (PLA2) activity in a concentration- and time-dependent manner. Superoxide generation as well as arachidonic acid (AA) release evoked by treatment with WKYMVm could be almost completely blocked by pretreatment of the cells with cytosolic PLA2 (cPLA2)-specific inhibitors. The involvement of cPLA2 in the peptide-induced AA release was further supported by translocation of cPLA2 to the nuclear membrane of monocytes incubated with WKYMVm. WKYMVm-induced phosphatidylbutanol formation was completely abolished by pretreatment with PKC inhibitors. Immunoblot showed that monocytes express phospholipase D1 (PLD1), but not PLD2. GF109203X as well as butan-1-ol inhibited peptide-induced superoxide generation in monocytes. Furthermore, the interrelationship between the two phospholipases, cPLA2 and PLD1, and upstream signaling molecules involved in WKYMVm-dependent activation was investigated. The inhibition of cPLA2 did not blunt peptide-stimulated PLD1 activation or vice versa. Intracellular Ca2+ mobilization was indispensable for the activation of PLD1 as well as cPLA2. The WKYMVm-dependent stimulation of cPLA2 activity was partially dependent on the activation of PKC and mitogen-activated protein kinase, while PKC activation, but not mitogen-activated protein kinase activation, was an essential prerequisite for stimulation of PLD1. Taken together, activation of the two phospholipases, which are absolutely required for superoxide generation, takes place through independent signaling pathways that diverge from a common pathway at a point downstream of Ca2+.  相似文献   

17.
The current study examined the signal transduction steps involved in the selective release of arachidonic acid (AA) induced by the addition of secretory phospholipase A2 (sPLA2) isotypes to bone marrow-derived mast cells (BMMC). Overexpression of sPLA2 receptors caused a marked increase in AA and PGD2 release after stimulation of BMMC, implicating sPLA2 receptors in this process. The hypothesis that the release of AA by sPLA2 involved activation of cytosolic PLA2 (cPLA2) was next tested. Addition of group IB PLA2 to BMMC caused a transient increase in cPLA2 activity and translocation of this activity to membrane fractions. Western analyses revealed that these changes in cPLA2 were accompanied by a time-dependent gel shift of cPLA2 induced by phosphorylation of cPLA2 at various sites. A noncatalytic ligand of the sPLA2 receptor, p-amino-phenyl-alpha-D-mannopyranoside BSA, also induced an increase in cPLA2 activity in BMMC. sPLA2 receptor ligands induced the phosphorylation of p44/p42 mitogen-activated protein kinase. Additionally, an inhibitor of p44/p42 mitogen-activated protein kinase (PD98059) significantly inhibited sPLA2-induced cPLA2 activation and AA release. sPLA2 receptor ligands also increased Ras activation while an inhibitor of tyrosine phosphorylation (herbimycin) inhibited the increase in cPLA2 activation and AA release. Addition of partially purified sPLA2 from BMMC enhanced cPLA2 activity and AA release. Similarly, overexpression of mouse groups IIA or V PLA2 in BMMC induced an increase in AA release. These data suggest that sPLA2 mediate the selective release of AA by binding to cell surface receptors and then inducing signal transduction events that lead to cPLA2 activation.  相似文献   

18.
Stimulation of rat mesangial cells for 24 h with interleukin-1beta (IL- 1beta) plus forskolin (Fk) leads to a marked increase in prostaglandin E2 (PGE2) synthesis. This effect is further enhanced by the small G-protein Rho inhibitor toxin A. A similar increase in PGE2 formation is obtained with Y27632, a Rho-dependent kinase inhibitor, and with lovastatin, a hydroxymethylglutaryl-coenzyme A inhibitor which depletes cells from geranylgeranyl moieties and thus blocks Rho activation. In parallel to the increased PGE2 synthesis, a potentiation of IL-1beta-induced secretory group IIA phospholipases A2 (sPLA2-IIA) protein expression also occurs by Rho inhibition. However, only toxin A triggers an increased sPLA2-IIA activity consistent with the elevated levels of protein expression, whereas Y27632 and lovastatin rather reduced IL-1beta-induced sPLA2-IIA activity. In vitro activity studies reveal that Y27632 and lovastatin can directly block sPLA2-IIA enzyme activity in a concentration-dependent manner. Interestingly, in the absence of IL-1beta/Fk stimulation and the lack of sPLA2-IIA protein expression, all Rho inhibitors exert a small but significant increase in PGE2 formation suggesting that additional PLA2s or downstream enzymes like cyclooxygenases or prostaglandin synthases may be activated by Rho inhibitors. Western blot analyses of toxin A-, Y27632- and lovastatin-stimulated cells reveal that the cytosolic group IV PLA2 (cPLA2) and the cytosolic PGE2 synthase (cPGES), but not the sPLA2-IIA, cyclooxygenase-2 or the microsomal PGE2 synthase (mPGES), are upregulated compared to unstimulated cells. Furthermore, the Rho inhibitors induced arachidonic acid release from intact cells which is blocked by the cPLA2 inhibitor methyl arachidonyl fluorophosphonate (MAFP). In summary, these data show that inhibition of the small G-protein Rho, either by toxin A, lovastatin, or Y27632, exert a dual effect on mesangial cells: (i) in the absence of an inflammatory stimulus it activates the constitutive cPLA2 and cPGE2 synthase and generates low amount of PGE2. (ii) In the presence of inflammatory cytokines it potentiates sPLA2-IIA expression and subsequent PGE2 formation. In addition, we identified lovastatin and Y27632 as direct inhibitors of sPLA2-IIA in a cell-free system.  相似文献   

19.
Whereas exogenous types IB and X secretory phospholipase A(2) (sPLA(2)) elicited prostaglandin D(2) (PGD(2)) production in mouse bone marrow-derived mast cells (BMMC), sPLA(2)-IIA was unable to do so. In search of a mechanism underlying this cellular refractoriness to exogenous sPLA(2)-IIA, we now report that this isozyme is promptly associated with cell surfaces, internalized, and then degraded in BMMC. Adsorption of sPLA(2)-IIA to BMMC was prevented by addition of heparin to the medium. Moreover, a heparin-nonbinding sPLA(2)-IIA mutant did not bind to BMMC. These results indicate that this sPLA(2)-IIA inactivation process depends on its rapid binding to heparan sulfate proteoglycan (HSPG) on BMMC surfaces. Thus, the present observations represent a particular situation in which cell surface HSPG exhibit a negative regulatory effect on cellular function of sPLA(2)-IIA, and argue that HSPG does not always act as a functional adapter for heparin-binding sPLA(2)s in mammalian cells as has been demonstrated before.  相似文献   

20.
The first step in prostacyclin (PGI(2)) synthesis involves the generation of arachidonic acid (AA) from membrane phospholipids mediated by the 85 kDa cytosolic phospholipase A(2) (cPLA(2)alpha). The current study examined the effects of secretory PLA(2)s (sPLA(2)s) on PGI(2) production by human umbilical vein endothelial cells (HUVEC). We demonstrate that exposure of HUVEC to sPLA(2) dose- and time-dependently enhances AA release and PGI(2) generation. sPLA(2)-stimulated AA mobilisation was blocked by AACOCF(3), an inhibitor of cPLA(2)alpha, suggesting cross-talk between the two classes of PLA(2). sPLA(2) induced the phosphorylation of cPLA(2)alpha and enhanced the phosphorylation states of p42/44(mapk), p38(mapk), and JNK, concomitant with elevated AA and PGI(2) release. The MEK inhibitor PD98059 attenuated sPLA(2)-stimulated cPLA(2)alpha phosphorylation and PGI(2) release. These data show that sPLA(2) cooperates with cPLA(2)alpha in a MAPK-dependent manner to regulate PGI(2) generation and suggests that cross-talk between sPLA(2) and cPLA(2)alpha is a physiologically important mechanism for enhancing prostanoid production in endothelial 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