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 156 毫秒
1.
Han J  Shin I 《Cellular signalling》2000,12(11-12):731-736
We have investigated the roles of ceramide in Fas signalling leading to phospholipase D (PLD) activation in A20 cells. Upon stimulation of Fas signalling by anti-Fas monoclonal antibody, sphingomyelin hydrolysis and activation of PLD were induced. Also, the translocation of protein kinase C (PKC) betaI and betaII and the elevation of diacylglycerol (DAG) content were induced by Fas cross-linking. When phosphatidylcholine-specific phospholipase C (PC-PLC) was inhibited by D609, the Fas-induced changes in PLD activity, DAG content, and PKC translocation were inhibited. In contrast, D609 had no effect on Fas-induced alterations in sphingolipid metabolism, suggesting that changes in ceramide content do not account for Fas-induced PLD activation. Furthermore, C6-ceramide had no effect on Fas-induced PLD activation and PKC translocation. Taken together, these data might suggest that ceramide generated by Fas cross-linking does not affect PKC beta-dependent PLD activity stimulated by anti-Fas monoclonal antibody in A20 cells.  相似文献   

2.
We have investigated the roles of ceramide in Fas signalling leading to phospholipase D (PLD) activation in A20 cells. Upon stimulation of Fas signalling by anti-Fas monoclonal antibody, sphingomyelin hydrolysis and activation of PLD were induced. Also, the translocation of protein kinase C (PKC) βI and βII and the elevation of diacylglycerol (DAG) content were induced by Fas cross-linking. When phosphatidylcholine-specific phospholipase C (PC-PLC) was inhibited by D609, the Fas-induced changes in PLD activity, DAG content, and PKC translocation were inhibited. In contrast, D609 had no effect on Fas-induced alterations in sphingolipid metabolism, suggesting that changes in ceramide content do not account for Fas-induced PLD activation. Furthermore, C6-ceramide had no effect on Fas-induced PLD activation and PKC translocation. Taken together, these data might suggest that ceramide generated by Fas cross-linking does not affect PKC β-dependent PLD activity stimulated by anti-Fas monoclonal antibody in A20 cells.  相似文献   

3.
The activation of phospholipase D in murine B cell lymphoma A20 cells treated with anti-Fas monoclonal antibody has been investigated. Fas cross-linking resulted in a both dose- and time-dependent increases in phospholipase D activity. There was a nearly maximum saturated rise in phospholipase D activity at the dose of 200 ng/ml anti-Fas monoclonal antibody showing a fourfold increase within 3 h. Fas activation also caused an approximately twofold increase of phosphatidylcholine-specific phospholipase C activity and 1,2-diacylglycerol release, which could be blocked by 30 min pretreatment with the phosphatidylcholine-specific phospholipase C inhibitor D609 (50 microgram/ml). Pretreatment of D609 also effectively inhibited the translocation of protein kinase C betaI and betaII from the cytosol to the membrane and the activation of phospholipase D induced by Fas cross-linking, suggesting that 1, 2-diacylglycerol released from the cellular phosphatidylcholine pool through phosphatidylcholine-specific phospholipase C plays a major role in protein kinase C/phospholipase D activation. Anti-Fas monoclonal antibody failed to elicit phosphoinositide-specific phospholipase C activation and any changes in the intracellular Ca2+ level in A20 cells, indicating that the phosphoinositide-mediated pathway is not involved in this Fas signaling. Therefore, these results suggest that Fas-mediated phospholipase D activation may be a consequence of primary stimulation of phosphatidylcholine-specific phospholipase C and that phospholipase D may play a role in Fas cross-linking signaling downstream from phosphatidylcholine-specific phospholipase C.  相似文献   

4.
We have previously reported that Fas cross-linking resulted in an increase in phospholipase D activity in A20 murine cells (J.-S. Han et al., Arch. Biochem. Biophys. 367, 233-239, 1999). In an attempt to explore the Fas downstream factor contributing to the activation of phospholipase D, we have investigated the possible involvement of a small GTP biding protein Ras in signaling events that were triggered by Fas cross-linking. Upon adenoviral expression of dominant negative mutant of Ras (N17Ras), an increase in phospholipase D activity by anti-Fas monoclonal antibody was diminished. Also, the Fas downstream signaling events triggered by Fas cross-linking such as the activation of phosphatidylcholine-specific phospholipase C, the increase in diacylglycerol level, and the translocation of protein kinase C to membrane fraction were all reduced by N17Ras expression. When parallel experiments were performed with manumycin-A, a Ras farnensyltransferase inhibitor, almost identical inhibitory effects on Fas downstream signaling were exhibited. These data suggest that Ras GTPase is essential in transmitting phospholipase D activation signal induced by Fas cross-linking and is located at phosphatidylcholine-specific phospholipase C upstream in Fas signaling cascades.  相似文献   

5.
The early signals generated following cross-linking of Fas/APO-1, a transmembrane receptor whose engagement by ligand results in apoptosis induction, were investigated in human HuT78 lymphoma cells. Fas/APO-1 cross-linking by mAbs resulted in membrane sphingomyelin hydrolysis and ceramide generation by the action of both neutral and acidic sphingomyelinases. Activation of a phosphatidylcholine-specific phospholipase C (PC-PLC) was also detected which appeared to be a requirement for subsequent acidic sphingomyelinase (aSMase) activation, since PC-PLC inhibitor D609 blocked Fas/APO-1-induced aSMase activation, but not Fas/APO-1-induced neutral sphingomyelinase (nSMase) activation. Fas/APO-1 cross-linking resulted also in ERK-2 activation and in phospholipase A2 (PLA2) induction, independently of the PC-PLC/aSMase pathway. Evidence for the existence of a pathway directly involved in apoptosis was obtained by selecting HuT78 mutant clones spontaneously expressing a newly identified death domain-defective Fas/APO-1 splice isoform which blocks Fas/APO-1 apoptotic signalling in a dominant negative fashion. Fas/APO-1 cross-linking in these clones fails to activate PC-PLC and aSMase, while nSMase, ERK-2 and PLA2 activates are induced. These results strongly suggest that a PC-PLC/aSMase pathway contributes directly to the propagation of Fas/APO-1-generated apoptotic signal in lymphoid cells.  相似文献   

6.
Kim JH  Yoon YD  Shin I  Han JS 《IUBMB life》1999,48(4):445-452
Although recent studies have demonstrated that ovarian follicular atresia occurs by apoptosis of granulosa cells, the intracellular signaling pathways involved in apoptotic cell death are still poorly characterized. We examined the role of ceramide as a candidate intracellular mediator of Fas-mediated signaling in cultured granulosa cells. Expression of Fas antigen was demonstrated by Western blot of granulosa cell lysates and immunostaining of cultured granulosa cells. Exposure of granulosa cells to anti-Fas monoclonal antibody (anti-Fas mAb) resulted in significant sphingomyelin hydrolysis, which was accompanied by a progressive increase in endogenous levels of ceramide. The addition of exogenous C6-ceramide induced drastic morphological change, including nuclear fragmentation and typical apoptotic DNA degradation. Furthermore, both anti-Fas mAb and C6-ceramide decreased phospholipase D (PLD) activity and diacylglycerol (DAG) concentrations in a time- or a dose-dependent manner. In addition, treatment with phorbol 12-myristate 13-acetate completely attenuated the ceramide-induced inhibition of PLD activity and partially suppressed ceramide-induced apoptosis. These results indicate that the Fas/ceramide signaling pathway might play a role in granulosa cell apoptosis and suggest that the PLD/DAG pathway might be cross-linked to the Fas/ceramide pathway in apoptotic processes of granulosa cells.  相似文献   

7.
In order to evaluate the possible contribution of phospholipase D (PLD) stimulation to the mitogenic response, a screening of a variety of different compounds, some of which are known to be potent mitogens, was performed using the well characterized Chinese hamster lung fibroblast (CCL39) cell line. In wild type CCL39 cells, or derivatives expressing high levels of either the human M1 muscarinic receptor (Hm1) or the human epidermal growth factor (EGF) receptor (39M1-81 and 39ER22 clones, respectively), thrombin, a potent mitogen for all three cell types, elicited the rapid activation of PLD (t1/2 activation, 30 s). Carbachol-mediated activation of the Hm1 receptor in the 39M1-81 clone, which is not a mitogenic signal, produced a similarly rapid although greater activation of PLD. Addition of EGF to the 39ER22 clone was able to provoke both a mitogenic response and stimulate PLD, albeit a comparatively small effect. In each case, the stimulation of PLD correlated closely with the ability to stimulate inositol phospholipid breakdown and was entirely dependent on the activation of protein kinase C. Moreover, the ability of both thrombin and carbachol to stimulate PLD was found to be rapidly desensitized, with a similar time course of desensitization (t1/2 desensitization, 90 s). It has recently been reported that an increase in phospholipase C (PLC)-mediated phosphocholine (PC) hydrolysis by either addition of agonist or by extracellular addition of PC-specific PLC enzyme constitutes a mitogenic signal. In this regard, in addition to stimulation of PLD, thrombin and carbachol were both able to stimulate the activity of a phosphocholine-specific phospholipase C (PC-PLC), which did not appear to desensitize within the time course employed. By contrast, EGF was unable to elicit the stimulation of PC-PLC. Ligands such as fibroblast growth factor (FGF) and platelet-derived growth factor (PDGF), which bind to and activate receptors with intrinsic tyrosine kinase activity, are potent mitogens for CCL39 cells but were unable to stimulate either PLD or PC-PLC activity. Furthermore, exogenous addition of purified PC-PLC enzyme, although able to induce a strong and lasting hydrolysis of PC, was unable to produce a mitogenic signal on its own. On the basis of these results, we conclude that the activation of both PLD and PC-PLC is neither sufficient nor required to produce a mitogenic response.  相似文献   

8.
The CD38 cell surface receptor is a potent activator for splenic, B lymphocytes. The molecular mechanisms regulating this response, however, remain incompletely characterized. Activation of the nonreceptor tyrosine kinase, Btk, is essential for CD38 downstream signaling function. The major Btk-dependent substrate in B cells, phospholipase C-gamma2 (PLC-gamma2), functions to generate the key secondary messengers, inositol-1,4,5 trisphosphate and diacylglycerol. Surprisingly, CD38 ligation results in no detectable increase in phosphoinositide metabolism and only a minimal increase in cytosolic calcium. We hypothesized that Btk functioned independently of PLC-gamma2 in the CD38 signaling pathway. Accordingly, we demonstrate that CD38 cross-linking does not result in the functional phosphorylation of PLC-gamma2 nor an increase in inositol-1,4,5 trisphosphate production. Furthermore, splenic B cells exhibit a normal CD38-mediated, proliferative response in the presence of the phosphoinositide-PLC inhibitor, U73122. Conversely, protein kinase C (PKC) beta-deficient mice, or PKC inhibitors, indicated the requirement for diacylglycerol-dependent PKC isoforms in this pathway. Loss of PKC activity blocked CD38-dependent, B cell proliferation, NF-kappaB activation, and subsequent expression of cyclin-D2. These results suggested that an alternate diacylglycerol-producing phospholipase must participate in CD38 signaling. Consistent with this idea, CD38 increased the enzymatic activity of the phosphatidylcholine (PC)-metabolizing enzymes, PC-PLC and phospholipase D. The PC-PLC inhibitor, D609, completely blocked CD38-dependent B cell proliferation, IkappaB-alpha degradation, and cyclin-D2 expression. Analysis of Btk mutant B cells demonstrated a partial requirement for Btk in the activation of both enzymes. Taken together, these data demonstrate that CD38 initiates a novel signaling cascade leading to Btk-, PC-PLC-, and phospholipase D-dependent, PLC-gamma2-independent, B lymphocyte activation.  相似文献   

9.
Many studies have shown that protein kinase C (PKC) is an important physiological regulator of phospholipase D (PLD). However, the role of PKC in agonist-induced PLD activation has been mainly investigated with a focus on the PLD1, which is one of the two PLD isoenzymes (PLD1 and PLD2) cloned to date. Since the expression of PLD2 significantly enhanced phorbol 12-myristate 13-acetate (PMA)- or bradykinin-induced PLD activity in rat pheochromocytoma PC12 cells, we investigated the regulatory mechanism of PLD2 in PC12 cells. Two different PKC inhibitors, GF109203X and Ro-31-8220, completely blocked PMA-induced PLD2 activation. In addition, specific inhibition of PKC delta by rottlerin prevented PLD2 activation in PMA-stimulated PC12 cells. Concomitant with PLD2 activation, PLD2 became phosphorylated upon PMA or bradykinin treatment of PC12 cells. Moreover, rottlerin blocked PMA- or bradykinin-induced PLD2 phosphorylation in PC12 cells. Expression of a kinase-deficient mutant of PKC delta using adenovirus-mediated gene transfer inhibited the phosphorylation and activation of PLD2 induced by PMA in PC12 cells, suggesting the phosphorylation-dependent regulation of PLD2 mediated by PKC delta kinase activity in PC12 cells. PKC delta co-immunoprecipitated with PLD2 from PC12 cell extracts, and associated with PLD2 in vitro in a PMA-dependent manner. Phospho-PLD2 immunoprecipitated from PMA-treated PC12 cells and PLD2 phosphorylated in vitro by PKC delta were resolved by two-dimensional phosphopeptide mapping and compared. At least seven phosphopeptides co-migrated, indicating the direct phosphorylation of PLD2 by PKC delta inside the cells. Immunocytochemical studies of PC12 cells revealed that after treatment with PMA, PKC delta was translocated from the cytosol to the plasma membrane where PLD2 is mainly localized. These results suggest that PKC delta-dependent direct phosphorylation plays an important role in the regulation of PLD2 activity in PC12 cells.  相似文献   

10.
This study uses human alveolar macrophages to determine whether activation of a phosphatidylcholine (PC)-specific phospholipase C (PC-PLC) is linked to activation of the p42/44 (ERK) kinases by LPS. LPS-induced ERK kinase activation was inhibited by tricyclodecan-9-yl xanthogenate (D609), a relatively specific inhibitor of PC-PLC. LPS also increased amounts of diacylglycerol (DAG), and this increase in DAG was inhibited by D609. LPS induction of DAG was, at least in part, derived from PC hydrolysis. Ceramide was also increased in LPS-treated alveolar macrophages, and this increase in ceramide was inhibited by D609. Addition of exogenous C2 ceramide or bacterial-derived sphingomyelinase to alveolar macrophages increased ERK kinase activity. LPS also activated PKC zeta, and this activation was inhibited by D609. LPS-activated PKC zeta phosphorylated MAP kinase kinase, the kinase directly upstream of the ERK kinases. LPS-induced cytokine production (RNA and protein) was also inhibited by D609. As an aggregate, these studies support the hypothesis that one way by which LPS activates the ERK kinases is via activation of PC-PLC and that activation of a PC-PLC is an important component of macrophage activation by LPS.  相似文献   

11.
The regulation of phospholipase D1 (PLD1), which has been shown to be activated by protein kinase C (PKC) alpha, was investigated in the human melanoma cell lines. In G361 cell line, which lacks PKCalpha, 12-O-tetradecanoylphorbol-13-acetate (TPA)-induced PLD activation was potentiated by introducing PKCalpha by the adenovirus vector. The kinase-negative PKCalpha elevated TPA-induced PLD activity less significantly than the wild type. A PKC specific inhibitor GF109203X lowered PLD activation in the cells expressing PKCalpha, but did not prevent PLD potentiation induced by the kinase-negative PKCalpha. Expression of PKCbetaII and the kinase-negative PKCbetaII enhanced TPA-stimulated PLD activity moderately in MeWo cell line, in which PKCbetaII is absent. Furthermore, the TPA treatment increased the association of PKCalpha, PKCbetaII, and their kinase-negative mutants with PLD1 in melanoma cells. These results indicate that PLD1 is dually regulated through phosphorylation as well as through the protein-protein interaction by PKCalpha, and probably by PKCbetaII, in vivo.  相似文献   

12.
1alpha,25-(OH)(2)D(3) regulates protein kinase C (PKC) activity in growth zone chondrocytes by stimulating increased phosphatidylinositol-specific phospholipase C (PI-PLC) activity and subsequent production of diacylglycerol (DAG). In contrast, 24R,25-(OH)(2)D(3) regulates PKC activity in resting zone (RC) cells, but PLC does not appear to be involved, suggesting that phospholipase D (PLD) may play a role in DAG production. In the present study, we examined the role of PLD in the physiological response of RC cells to 24R,25-(OH)(2)D(3) and determined the role of phospholipases D, C, and A(2) as well as G-proteins in mediating the effects of vitamin D(3) metabolites on PKC activity in RC and GC cells. Inhibition of PLD with wortmannin or EDS caused a dose-dependent inhibition of basal [3H]-thymidine incorporation by RC cells and further increased the inhibitory effect of 24R,25-(OH)(2)D(3). Wortmannin also inhibited basal alkaline phosphatase activity and [35]-sulfate incorporation and decreased the stimulatory effect of 24R,25-(OH)(2)D(3). This inhibitory effect of wortmannin was not seen in cultures treated with the PI-3-kinase inhibitor LY294002, verifying that wortmannin affected PLD. Wortmannin also inhibited basal PKC activity and partially blocked the stimulatory effect of 24R,25-(OH)(2)D(3) on this enzyme activity. Neither inhibition of PI-PLC with U73122, nor PC-PLC with D609, modulated PKC activity. Wortmannin had no effect on basal PLD in GC cells, nor on 1alpha,25-(OH)(2)D(3)-dependent PKC. Inhibition of PI-PLC blocked the 1alpha,25-(OH)(2)D(3)-dependent increase in PKC activity but inhibition of PC-PLC had no effect. Activation of PLA(2) with melittin inhibited basal and 24R,25-(OH)(2)D(3)-stimulated PKC in RC cells and stimulated basal and 1alpha,25-(OH)(2)D(3)-stimulated PKC in GC cells, but wortmannin had no effect on the melittin-induced changes in either cell type. Pertussis toxin modestly increased the effect of 24R,25-(OH)(2)D(3) on PKC, whereas GDPbetaS had no effect, suggesting that PLD2 is the isoform responsible. This indicates that 1alpha,25-(OH)(2)D(3) regulates PKC in GC cells via PI-PLC and PLA(2), but not PC-PLC or PLD, whereas 24R,25-(OH)(2)D(3) regulates PKC in RC cells via PLD2.  相似文献   

13.
A 66-kDa molecular weight protein with phospholipase D activity was solubilized and partially purified from rat liver plasma membrane. The activity and regulation of this phospholipase D have been characterized. Immunoblot analyses indicated that the enzyme was distinct from hPLD1 and PLD2, but was recognized by an antibody to the 12 terminal amino acids of PLD1. PLD activity was stimulated by 1-100 microM Ca(2+) and Mg(2+) and displayed a pH optimum of 7.5. Activity was inhibited by both saturated and unsaturated fatty acids. This PLD was activated in an ATP-independent manner by the PKC isozymes alpha and betaII but not activated by other PKC isozymes. It was also stimulated by the small G-proteins RhoA and ARF. RhoA stimulated the greatest activation, followed by ARF and PKC(alpha). This enzyme was further activated in a synergistic manner when combinations of PKC(alpha) and RhoA or ARF were used. This enzyme displayed a greater response activation by RhoA than to activation by ARF. While a potential breakdown product of PLD1, activation by RhoA indicates that the PLD characterized here is distinct from the other PLDs cloned or isolated to date.  相似文献   

14.
Mammalian spermatozoa should reside in the female reproductive tract for a certain time before gaining the ability to fertilize. During this time, the spermatozoa undergo a series of biochemical processes collectively called capacitation. We recently demonstrated that actin polymerization is a necessary step in the cascade leading to capacitation. We demonstrate here for the first time a role for phospholipase D (PLD) in the induction of actin polymerization and capacitation in spermatozoa. The involvement of PLD is supported by specific inhibition of F-actin formation during sperm capacitation by PLD inhibitors and the stimulation of fast F-actin formation by exogenous PLD or phosphatidic acid (PA). Moreover, PLD activity is enhanced during capacitation before actin polymerization. Protein kinase A (PKA), known to be active in sperm capacitation, and protein kinase C (PKC), involved in the acrosome reaction, can both activate PLD and actin polymerization. We suggest that PKA- and PKC-dependent signal transduction pathways can potentially lead to PLD activation; however, under physiological conditions, actin polymerization depends primarily on PKA activity. Activation of PKA during capacitation causes inactivation of phospholipase C, and as a result, PKC activation is prevented. It appears that PKA activation promotes sperm capacitation whereas early activation of PKC during capacitation would jeopardize this process.  相似文献   

15.
Murine B lymphoma A20 cells are highly sensitive to Fas-mediated death signals induced by anti-Fas antibody Jo2 or cross-linked Fas ligand (FasL). We have found that the microfilament poison cytochalasin D blocks Fas-mediated apoptosis induced by Jo2 but not FasL in A20 cells. The induction of Fas-mediated apoptosis by Jo2 was antagonized by anti-Fcgamma RII/RIII receptor (FcgammaR) antibody, and defective in FcgammaR-negative A20 cells. Since the induction of Jo2-mediated apoptosis in FcgammaR-negative A20 cells was reversed by the addition of wild type A20 cells or the cross-linking agent protein A, Fas-expressing bystander A20 cells seem to be killed by other A20 cells that capture and cross-link monomeric Jo2 via FcgammaR. Although cytochalasin D affected FcgammaR-mediated cross-linking of Jo2 molecules, the drug markedly inhibited the intracellular signaling pathway induced by Jo2. The blockade of Jo2-induced apoptosis by cytochalasin D occurred upstream of caspase-8 activation. Thus, these observations suggest that actin cytoskeleton is required for early apoptosis signaling induced by Jo2, but not physiological FasL.  相似文献   

16.
Regulation of phospholipase D (PLD) activity participating in signal transduction involves complex interactions with small G-proteins (ARF, Rho) and protein kinase C isoforms (PKCalpha). In SK-N-MC human neuroblastoma cells, phorbol ester (TPA) activation of PLD was enhanced by overexpressing myristoylated alanine-rich C kinase substrate (MARCKS). To study MARCKS interactions with PLD, we investigated PLD isoform expression and activation by TPA and GTPgammaS in intact and digitonin-permeabilized clones transfected with MARCKS (M22). PLD2 was in both cytosol and membrane fractions while PLD1 was primarily membrane-associated in both vector control and M22 cells; location or quantities were unaltered by TPA treatment. TPA-stimulated PLD activity was higher in both intact and digitonin-permeabilized M22 cells than in vector controls. In contrast, GTPgammaS-stimulated PLD activity was independent of MARCKS expression but was additive with MARCKS-PKC-dependent activation in permeabilized cells. Combinations of PKC inhibition and down-regulation in intact and permeabilized (with GTPgammaS present) cells indicated that a PKC-mediated phosphorylation event was necessary in intact cells without access to GTPgammaS, stimulation of PLD mediated by GTPgammaS was independent of PKC, and PLD activation by PKC in permeabilized cells was kinase-independent. Western blot analysis showed that MARCKS, PKCalpha, PLD1 and PLD2 were present in a detergent-insoluble fraction (DIF); GTPgammaS increased recovery of PLD2 in DIF. Disruption of cholesterol-rich DIFs with digitonin, cyclodextrin or filipin potentiated activation of PLD by TPA. Our studies suggest that activation of PLD by PKC requires MARCKS and can involve both phosphorylation-independent and -dependent processes. As PLD activation by GTPgammaS is PKC-MARCKS-independent, MARCKS may provide a fine tuning component in conjunction with G-protein-mediated mechanisms for regulation of PLD.  相似文献   

17.
The Fas antigen is a member of the tumor necrosis factor/nerve growth factor receptor family and is expressed in tissues such as the thymus, liver, and ovary. Agonistic anti-Fas antibodies have cytolytic activity against cell lines expressing the Fas antigen. In this study, we show that anti-Fas antibody can induce the death of mouse hepatocytes in primary culture. Cell death via apoptosis was evidenced by the fact that the dying cells displayed DNA fragmentation, extensive surface bleb formation, and chromatin condensation. Anti-Fas antibody alone induced apoptosis in less than 20% of the cultured hepatocytes, whereas all cells were killed within 24 h by anti-Fas antibody in the presence of actinomycin D, cycloheximide, or protein kinase C (PKC) inhibitors such as H-7 and HA1004. These results indicate that the Fas antigen expressed in mouse hepatocytes functionally transduces the apoptotic signal and suggest that cultured mouse hepatocytes express protective proteins against apoptosis and that phosphorylation by PKC is also involved in protection of the hepatocytes from Fas-mediated apoptosis.  相似文献   

18.
Several factors for the hydrogen peroxide (H(2)O(2))-induced PLD stimulation have been proposed, including protein kinase C (PKC), tyrosine kinase, mitogen-activated protein kinase and Ca(2+), but their precise roles remain to be defined. As for involvement of PKC, there has been some discrepancy. Our previous study has demonstrated that phospholipase D (PLD) activity was increased by exposure of PC12 cells to 0.5mM H(2)O(2) in modified Krebs-Ringer buffer (KRB) and suggested that the PLD activation was independent of PKC activity. However, we have shown here that the H(2)O(2)-induced PLD stimulation was much greatly enhanced by incubation in Dulbecco's modified Eagle's medium (DMEM) and further that it was PKC-dependent. These results indicated that the markedly enhanced PLD activation and its PKC dependence were modulated by pH changes during incubation in DMEM. Furthermore, evidence has been presented for possible involvement of alkaline phosphatase in this pH-dependent profile of PLD activation by H(2)O(2).  相似文献   

19.
We found that a short synthetic peptide corresponding to the antisense homology box of Fas ligand induced apoptotic cell death of Fas-expressing human ovarian cancer cell lines. The peptide was deduced from residues 256–265 of human Fas ligand, based on the hypothesis that it should contain a specific binding site to the corresponding Fas. Interestingly, the ovarian cancer cell line NOS4, which was sensitive to anti-Fas antibody induced apoptosis, was not affected by the peptide, whereas another cell line, SKOV-3, which was insensitive to anti-Fas antibody, was killed by the peptide. Thus, this short peptide was shown to have a unique activity to induce apoptosis in human ovarian cancer cells in a manner different from anti-Fas antibody.  相似文献   

20.
In inflammatory cells, agonist-stimulated arachidonic acid (AA) release is thought to be induced by activation of group IV Ca(2+)-dependent cytosolic phospholipase A(2) (cPLA(2)) through mitogen-activated protein kinase (MAP kinase)- and/or protein kinase C (PKC)-mediated phosphorylation and Ca(2+)-dependent translocation of the enzyme to the membrane. Here we investigated the role of phospholipases in N-formylmethionyl-l-leucyl-l-phenylalanine (fMLP; 1 nM-10 microM)-induced AA release from neutrophil-like db-cAMP-differentiated HL-60 cells. U 73122 (1 microM), an inhibitor of phosphatidyl-inositol-4,5-biphosphate-specific phospholipase C, or the membrane-permeant Ca(2+)-chelator 1, 2-bis?2-aminophenoxy?thane-N,N,N',N'-tetraacetic acid (10 microM) abolished fMLP-mediated Ca(2+) signaling, but had no effect on fMLP-induced AA release. The protein kinase C-inhibitor Ro 318220 (5 microM) or the inhibitor of cPLA(2) arachidonyl trifluoromethyl ketone (AACOCF(3); 10-30 microM) did not inhibit fMLP-induced AA release. In contrast, AA release was stimulated by the Ca(2+) ionophore A23187 (10 microM) plus the PKC activator phorbol myristate acetate (PMA) (0.2 microM). This effect was inhibited by either Ro 318220 or AACOCF(3). Accordingly, a translocation of cPLA(2) from the cytosol to the membrane fraction was observed with A23187 + PMA, but not with fMLP. fMLP-mediated AA release therefore appeared to be independent of Ca(2+) signaling and PKC and MAP kinase activation. However, fMLP-mediated AA release was reduced by approximately 45% by Clostridium difficile toxin B (10 ng/ml) or by 1-butanol; both block phospholipase D (PLD) activity. The inhibitor of phosphatidylcholine-specific phospholipase C (PC-PLC), D609 (100 microM), decreased fMLP-mediated AA release by approximately 35%. The effect of D609 + 1-butanol on fMLP-induced AA release was additive and of a magnitude similar to that of propranolol (0.2 mM), an inhibitor of phosphatidic acid phosphohydrolase. This suggests that the bulk of AA generated by fMLP stimulation of db-cAMP-differentiated HL-60 cells is independent of the cPLA(2) pathway, but may originate from activation of PC-PLC and PLD.  相似文献   

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