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 281 毫秒
1.
Phospholipase D (PLD) plays an important role as an effector in the membrane lipid-mediated signal transduction. However, the precise physiological functions of PLD are not yet well understood. In this study, we examined the role of PLD activity in hydrogen peroxide (H(2)O(2))-induced apoptosis in rat pheochromocytoma (PC12) cells. Treatment of PC12 cells with H(2)O(2) resulted in induction of apoptosis in these cells, which is accompanied by the activation of PLD. This H(2)O(2)-induced apoptosis was enhanced remarkably when phosphatidic acid production by PLD was selectively inhibited by pretreating the PC12 cells with 1-butanol. Expression of PLD2, but not of PLD1, correlated with increased H(2)O(2)-induced PLD activity in a concentration- and time-dependent manner. Concomitant with PLD activation, the PLD2 activity suppressed H(2)O(2)-induced apoptosis in PC12 cells. Expression of PLD2 lipase-inactive mutant (K758R) had no effect on either PLD activity or apoptosis. PLD2 activity also suppressed H(2)O(2)-induced cleavage and activation of caspase-3. Taken together, the results suggest that PLD2 activity is specifically up-regulated by H(2)O(2) in PC12 cells and that it plays a suppressive role in H(2)O(2)-induced apoptosis.  相似文献   

2.
Phospholipase D (PLD) is a ubiquitously expressed enzyme of ill-defined function. In order to explore its cellular actions, we inactivated the rat PLD1 (rPLD1) isozyme by tagging its C terminus with a V5 epitope (rPLD1-V5). This was stably expressed in Rat-2 fibroblasts to see if it acted as a dominant-negative mutant for PLD activity. Three clones that expressed rPLD1-V5 were selected (Rat2V16, Rat2V25, and Rat2V29). Another clone (Rat2V20) that lost expression of rPLD1-V5 was also obtained. In the three clones expressing rPLD1-V5, PLD activity stimulated by phorbol myristate acetate (PMA) or lysophosphatidic acid (LPA) was reduced by ~50%, while the PLD activity of Rat2V20 cells was normal. Changes in the actin cytoskeleton in response to LPA or PMA were examined in these clones. All three clones expressing rPLD1-V5 failed to form actin stress fibers after treatment with LPA. However, Rat2V20 cells formed stress fibers in response to LPA to the same extent as wild-type Rat-2 cells. In contrast, there was no significant change in membrane ruffling induced by PMA in the cells expressing rPLD1-V5. Since Rho is an activator both of rPLD1 and stress fiber formation, the activation of Rho was monitored in wild-type Rat-2 cells and Rat2V25 cells, but no significant difference was detected. The phosphorylation of vimentin mediated by Rho-kinase was also intact in Rat2V25 cells. Rat2V25 cells also showed normal vinculin-containing focal adhesions. However, the translocation of alpha-actinin to the cytoplasm and to the detergent-insoluble fraction in Rat2V25 cells was reduced. These results indicate that PLD activity is required for LPA-induced rearrangement of the actin cytoskeleton to form stress fibers and that PLD might be involved in the cross-linking of actin filaments mediated by alpha-actinin.  相似文献   

3.
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   

4.
Although the functional significance of neuronal phospholipase D (PLD) is being recognized, little is known about its regulatory role in neuronal cells. To elucidate the regulatory mechanism of neuronal PLD, we investigated PLD(2)-binding neuronal protein from rat brain cytosol. During the fractionation of rat brain cytosol by four-column chromatography, a 62-kDa PLD(2)-interacting protein was detected by PLD(2) overlay assay and identified as collapsin response mediator protein-2 (CRMP-2), which controls neuronal axon guidance and outgrowth. Using bacterially expressed glutathione S-transferase fusion proteins, we found that two regions (amino acids 65-192 (the phagocytic oxidase domain) and 724-825) of PLD(2) and a single region (amino acids 243-300) of CRMP-2 are required for the direct binding of both proteins. A co-immunoprecipitation study in COS-7 cells also showed an in vivo interaction between CRMP-2 and PLD(2). Interestingly, CRMP-2 was found to potently inhibit PLD(2) activity in a concentration-dependent manner (IC(50) = 30 nm). Overexpression studies also showed that CRMP-2 is an in vivo inhibitor of PLD(2) in PC12 cells. Moreover, increasing the concentration of semaphorin 3A, one of the repulsive axon guidance cues, showed that PLD(2) activity can be inhibited in PC12 cells. Immunocytochemistry further revealed that PLD(2) is co-localized with CRMP-2 in the distal tips of neurites, its possible action site, in differentiated PC12 cells. Taken together, our results indicate that CRMP-2 may interact directly with and inhibit neuronal PLD(2), suggesting that this inhibitory mode of regulation may play a role in neuronal pathfinding during the developmental stage.  相似文献   

5.
6.
Among the viral regulatory genes the tat and nef genes of HIV-1 encode the proteins playing a central role in viral replication and exerting pleiotropic effects on the survival and growth of the cells. These effects differ in various cell types, possibly due to the use of genes from different HIV-1 isolates. In this work, we studied the effects of the tat and nef genes on three types of cultured rat cells: primary embryo fibroblasts, pseudonormal Rat-2, and pheochromocytoma PC12. Both genes affected growth properties and morphology of cells, the effects being cell-specific. The proliferative activity of both Rat-2 and PC12 cells was considerably increased after transfection with the tat gene. In primary rat embryo fibroblasts the tat gene induced multilayered foci. More importantly, it was shown that the efficiency of transformation was higher in cells coexpressing tat and nef. The nef gene caused considerable suppression of Rat-2 cell proliferation, but no changes in their morphology. The nef gene transfection of PC12 cells also led to suppression of their proliferative activity. In addition, cellular agglomerates which were morphologically similar to multinuclear syncytial cells were detected in these cells for the first time.  相似文献   

7.
The role of phospholipase D (PLD) activation in hydrogen peroxide (H(2)O(2))-induced signal transduction and cellular responses is not completely understood. Here we present evidence that Ca(2+)-dependent tyrosine kinase, Pyk2, requires PLD activation to mediate survival pathways in rat pheochromocytoma PC12 cells under oxidative stress. The H(2)O(2)-induced phosphorylation of two Pyk2 sites (Tyr(580), and Tyr(881)) was suppressed by 1-butanol, an inhibitor of transphosphatidylation by PLD, and also by transfection of catalytically negative mouse PLD2K758R (PLD2KR). Furthermore, we found that PLD2 was associated with Pyk2 and Src, and that activation of PLD2 was required for H(2)O(2)-enhanced association of Src with Pyk2 leading to full activation of Pyk2. H(2)O(2)-induced phosphorylation of Akt and p70S6K was dependent on phosphatidylinositol 3-kinase (PI3K) activity and was abolished by 1-butanol but not t-butanol. Furthermore, the PI3K/Akt activation in response to H(2)O(2) was reduced by transfection of either PLD2KR or the dominant negative Pyk2DN. This study is the first demonstration that PLD2 activation is implicated in Src-dependent phosphorylation of Pyk2 (Tyr(580) and Tyr(881)) by promoting the complex formation between Pyk2 and activated Src in PC12 cells exposed to H(2)O(2), thereby resulting in activation of the survival signaling pathway PI3K/Akt/p70S6K.  相似文献   

8.
Cyclooxygenase-1 (COX-1) behaves as a delayed response gene in rat pheochromocytoma (PC12) cells exposed to nerve growth factor (NGF). To investigate the possible targets for COX-1 generated prostanoids in the early stages of neuronal differentiation, we have examined the expression of prostanoid receptors by PC12 cells using functional assays. Prostanoid receptor-specific agonists failed to activate adenylyl cyclase in undifferentiated and NGF-treated PC12 cells; neither did they stimulate phospholipase C activity. EP3 receptor agonists and PGF were the only active ligands, able to inhibit forskolin-stimulated adenylyl cyclase activity. PC12 cells expressed EP3 and FP receptor mRNA, but only the responses to EP3 receptor agonists were inhibited by the EP3 receptor antagonist ONO-AE3-240. The functional role of NGF-stimulated COX-1 remains to be determined since we found no strong evidence of a role for EP3 receptors in the morphological changes induced by NGF during the early stages of differentiation of PC12 cells.  相似文献   

9.
This paper presents evidence to suggest that interleukin-1 alpha as a complex with its receptor is able to express DNA binding activity. Both the interleukin-1/receptor complex and the interleukin-1 receptor appear to be able to bind to DNA, however interleukin-1 on its own showed no binding activity. Interleukin-1 was found to be internalised into the nuclei of all cells examined (EL4, MEL, HL-60, K562, THP-1 and Jurkat cells). The data suggest the possible modulation of genes by interaction of interleukin-1/receptor complexes with DNA structures.  相似文献   

10.
We studied the expression of the genes encoding the A and B chains of platelet-derived growth factor (PDGF) in a number of human leukemia cell lines. Steady-state expression of the A-chain RNA was seen only in the promonocytic leukemia cell line U937 and in the T-cell leukemia cell line MOLT-4. It has previously been reported that both PDGF A and PDGF B genes are induced during megakaryoblastic differentiation of the K562 erythroleukemia cells and transiently during monocytic differentiation of the promyelocytic leukemia cell line HL-60 and U937 cells. In this study we show that PDGF A RNA expression was induced in HL-60 and Jurkat T-cell leukemia cells and increased in U937 and MOLT-4 cells after a 1- to 2-h stimulation with an 8 pM concentration of transforming growth factor beta (TGF-beta). PDGF A RNA remained at a constant, elevated level for at least 24 h in U937 cells, but returned to undetectable levels within 12 h in HL-60 cells. No PDGF A expression was induced by TGF-beta in K562 cells or in lung carcinoma cells (A549). Interestingly, essentially no PDGF B-chain (c-sis proto-oncogene) RNA was expressed simultaneously with PDGF A. In the presence of TGF-beta and protein synthesis inhibitors, PDGF A RNA was superinduced at least 20-fold in the U937 and HL-60 cells. PDGF A expression was accompanied by secretion of immunoprecipitable PDGF to the culture medium of HL-60 and U937 cells. The phorbol ester tumor promoter tetradecanoyl phorbol acetate also increased PDGF A expression with similar kinetics, but with a mechanism distinct from that of TGF-beta. These results suggest a role for TGF-beta in the differential regulation of expression of the PDGF genes.  相似文献   

11.
Occupancy of chemotactic peptide receptors leads to rapid initiation of phospholipase D (PLD) activity in intact dimethylsulfoxide-differentiated HL-60 granulocytes (Pai, J.-K, Siegel, M.I., Egan, R.W., and Billah, M.M. (1988) J. Biol. Chem. 263, 12472). To gain further insight into the activation mechanisms, PLD has been studied in cell lysates from HL-60 granulocytes, using 1-0-alkyl-2-oleoyl-[32P]phosphatidylcholine (alkyl-[32P]PC), 1-0-[3H]alkyl-2-oleoyl-phosphatidylcholine [( 3H]alkyl-PC) and [14C]arachidonyl-phospholipids as substrates. In the presence of Ca2+ and GTP gamma S, post-nuclear homogenates degrade alkyl-[32P]PC to produce 1-0-alkyl-[32P]phosphatidic acid (alkyl-[32P]-PA), and in the presence of ethanol, also 1-0-alkyl-[32P]phosphatidylethanol (alkyl-[32P]PEt). By comparing the 3H/32P ratios of PA and PEt to that of PC, it is concluded that PA and PEt are formed exclusively by a PLD that catalyzes both hydrolysis and transphosphatidylation between PC and ethanol. Furthermore, PC containing either ester- or ether-linkage at the sn-1 position is degraded in preference to phosphatidylethanolamine and phosphatidylinositol by PLD in HL-60 cell homogenates. It is concluded that HL-60 granulocytes contain a PC-specific PLD that requires both Ca2+ and GTP for activation.  相似文献   

12.
Intact cells and cell-free systems were employed to characterize phospholipase D (PLD) activity in Madin-Darby canine kidney (MDCK) cells. In cells prelabeled with [3H]glycerol, 12-O-tetradecanoylphorbol-13-acetate (TPA) elicited phosphatidylcholine (PC) hydrolysis by PLD, as shown by the prolonged formation of [3H]phosphatidylethanol (PEt) and an accompanying decrease in [3H]PC. In contrast, bradykinin elicited rapid formation of [3H]PEt (approximately 1 min) accompanied by a decrease in [3H]phosphatidylinositol (PI). When the agonists were administered simultaneously, [3H]PEt formation was biphasic. In cells prelabeled with [3H] choline, at times less than 1 min, bradykinin failed to induce significant change in [3H]choline release. Bradykinin-induced formation of [3H]PEt in the [3H]glycerol-labeled cells was strictly dependent on extracellular Ca2+, whereas TPA-induced formation of [3H]PEt did not require extracellular Ca2+. Cell-free assays for PLD were used to assess the enzyme location, substrate specificity, and cofactor requirements. The PC-PLD activity (PEt formation) against [3H]stearoyl-PC was primarily localized in the 440 x g pellet (membrane- and nuclear-associated), preferred PC as a substrate, required detergent, and was not influenced by Ca2+ at low concentrations but was inhibited by Ca2+ in excess of 0.5 mM. The PI-PLD activity against [3H]stearoyl-PI was found largely in the 100,000 x g supernatant (cytosol), was strictly Ca(2+)-dependent, and did not require detergent. From these data, we conclude that MDCK cells contain two PLD subtypes: 1) a membrane-associated, PC-selective enzyme that responds to TPA resulting in prolonged hydrolysis of PC (the PC-PLD is Ca(2+)-independent, but requires detergent); 2) a cytosolic, PI-selective enzyme that responds rapidly but transiently to bradykinin (the PI-PLD requires Ca2+ but not detergent).  相似文献   

13.
It has recently been demonstrated that the chemotactic peptide N-formyl-Met-Leu-Phe activates phospholipase D (PLD) in dimethyl sulfoxide-differentiated HL-60 granulocytes to produce phosphatidic acid (PA) and, in the presence of ethanol, phosphatidylethanol (PEt) (Pai, J.-K., Siegel, M. I., Egan, R. W., and Billah, M. M. (1988) J. Biol. Chem. 263, 12472-12477). We now report that biologically active phorbol esters, a cell-permeable diacylglycerol, 1-oleoyl-2-acetylglycerol (OAG), and calcium ionophore A23187 are also potent inducers of PLD in these HL-60 granulocytes. HL-60 granulocytes have been selectively labeled in 1-O-alkyl-2-acyl-sn-glycero-3-phosphocholine (alkyl-PC) with 32P by incubating the cells with alkyl-[32P]lyso-phosphatidylcholine (PC). When these labeled cells are treated with phorbol 12-myristate 13-acetate (PMA), phorbol 12,13-dibutyrate, OAG, or A23187, alkyl-[32P]PA is formed. Because cellular ATP has not been labeled with 32P, the formation of alkyl-[32P]PA conclusively demonstrates PLD activation by these agents. In the presence of 0.5% ethanol, phorbol esters, OAG, and A23187 also induce formation of alkyl-[32P]PEt, demonstrating that the activated PLD catalyzes transphosphatidylation between the phosphatidyl moiety of the alkyl-[32P]PC and ethanol. Formation of alkyl-[32P]PA and alkyl-[32P]PEt in response to these various agents occurs in a time- and dose-dependent manner and exhibits differential Ca2+ requirements. Based on experiments with both [3H]alkyl-PC and alkyl-[32P]PC, it is concluded that alkyl-PA and alkyl-PEt formed in response to PMA, OAG, or A23187 are derived exclusively from PLD action on alkyl-PC. Furthermore, subthreshold concentrations of PMA (0.5-2.0 nM) or OAG (1.0-25 microM) combined with subthreshold levels of A23187 (15-60 nM) induce the formation of alkyl-[32P]PA and alkyl-[32P]PEt, suggesting that receptor-mediated activation of PLD might involve cooperative interactions between Ca2+ and diglyceride. Although PLD is activated by agents that also activate protein kinase C, the protein kinase C inhibitor, K252a, inhibits PMA-induced protein phosphorylation but causes only partial inhibition of PLD activation. We conclude that phorbol esters, OAG, and A23187 activate PLD in HL-60 granulocytes via protein kinase-independent as well as protein kinase-dependent mechanisms.  相似文献   

14.
Numerous studies show that PLD is activated in cells by calcium and by protein kinase C (PKC). We found that human PLD1 and PLD2 expressed in Sf9 cells can be activated by calcium-mobilizing agonists and by co-expression with PKCalpha. The calcium-mobilizing agonists A23187 and CryIC toxin triggered large increases in phosphatidylethanol (PtdEth) production in Sf9 cells over-expressing PLD1 and PLD2, but not in vector controls. PLD activation by these agonists was largely dependent on extracellular calcium. Membrane assays demonstrated significant PLD1 and PLD2 activity in the absence of divalent cations, which could be enhanced by low levels of calcium either in the presence or absence of magnesium. PLD1 but not PLD2 activity was slightly enhanced by magnesium. Treatment of Sf9 cells expressing PLD1 and PLD2 with PMA resulted in little PtdEth production. However, a significant and comparable formation of PtdEth occurred when PLD1 or PLD2 were co-expressed with PKCalpha, but not PKCdelta, and was further augmented by PMA. In contrast to PLD1, co-expressing PLD2 with PKCalpha or PKCdelta further enhanced A23187-induced PtdEth production. Immunoprecipitation experiments demonstrated that PLD1 and PLD2 associated with the PKC isoforms in Sf9 cells. Furthermore, in membrane reconstitution assays, both PLD1 and PLD2 could be stimulated by calmodulin and PKCalpha-enriched cytosol. The results indicate that PLD2 as well as PLD1 is subject to agonist-induced activation in intact cells and can be regulated by calcium and PKC.  相似文献   

15.
A series of 3-amino-1,2,4-benzotriazine-1,4-dioxide derivatives 1 have been synthesized and evaluated for their cytotoxic activity in vitro against human leukemia cell lines: Molt-4, K562, HL60, human liver cancer cell Hep-G2, human prostate cancer cell PC-3 in hypoxia. Most of the compounds showed more potent activity than TPZ. Compounds 1i and 1m displayed encouraging superior activity against Molt-4 and HL-60 cell lines. Three potential derivatives received the test of the activity in hypoxia and in normoxia against Molt-4 and HL-60 cell lines and showed obvious hypoxia selectivity. Further mechanism study revealed that the cytotoxic activities of compounds 1i and 1k in Molt-4 cells might be mediated by modulation of p53 protein expression and mitochondrial membrane potential (DeltaPsi(m)).  相似文献   

16.
The dose- and time-dependent antitumor and cytotoxic effects of L-asparaginases from Erwinia carotovora (ECAR LANS) and Escherichia coli (MEDAC) have been investigated using human leukemic cells and human and animal solid tumor cells. These included human T-cell acute lymphoblastic leukemia cell lines (Jurkat, Jurkat/A4, Molt-4), human chronic myeloid leukemia K562 cells, human promyelocytic leukemia HL-60, and also human solid tumor cells (prostate carcinoma LnCap, breast adenocarcinoma MCF7, ovarian adenocarcinoma SCOV-3 and carcinoma CaOV, hepatocarcinoma Hep G2, fibrosarcoma HT-1080) and animal solid tumor cells (rat Gasser??s ganglion neurinoma cells GGNC-1, mouse glioblastoma EPNT-5). We investigated sensitivity of tumor cells (seeded at different density) to L-asparaginases, as well the effect of L-asparaginases on cell growth rate, protein and DNA synthesis in the presence of various cytostatics. Cell cycle analysis by flow cytofluorimetry and detection of apoptotic cells before and after treatment with L-asparaginases indicate that ECAR LANS L-asparaginase suppressed growth of all tested solid tumor cells. Evaluation of leukemic cell number after treatment with L-asparaginases for 24, 48 and 72 h demonstrated that asparagine deficiency did not kill cells but stopped normal cell division. The cytofluorometric study of solid and leukemic cells revealed that except HL-60 cells the treatment with L-asparaginase for 72 h did not change cell cycle phase distribution and did not increase the number of apoptotic cells. Combined treatment of cells using a combination of L-asparaginase and doxorubicin significantly increased the number of apoptotic cells up to 60% (MCF-7 cells), 40% (Jurkat cells) and even 99% (HL-60). High levels of DNA and protein synthesis rates in asparaginase-treated tumor cells suggest lack of massive entry of tumor cells to apoptosis. This conclusion is based on the fact of sensitivity of multi-resistant Jurkat/A4 cells to L-asparaginases (it is nearly impossible to induce apoptosis in these cells). Since ECAR LANS did not influence growth of normal human fibroblasts it appears that the enzyme cytotoxicity is associated only asparagine deficiency.  相似文献   

17.
Cell differentiation is compromised in acute leukemias. We report that mammalian target of rapamycin (mTOR) and S6 kinase (S6K) are highly expressed in the undifferentiated promyelomonocytic leukemic HL-60 cell line, whereas PLD2 expression is minimal. The expression ratio of PLD2 to mTOR (or to S6K) is gradually inverted upon in vitro induction of differentiation toward the neutrophilic phenotype. We present three ways that profoundly affect the kinetics of differentiation as follows: (i) simultaneous overexpression of mTOR (or S6K), (ii) silencing of mTOR via dsRNA-mediated interference or inhibition with rapamycin, and (iii) PLD2 overexpression. The last two methods shortened the time required for differentiation. By determining how PLD2 participates in cell differentiation, we found that PLD2 interacts with and activates the oncogene Fes/Fps, a protein-tyrosine kinase known to be involved in myeloid cell development. Fes activity is elevated with PLD2 overexpression, phosphatidic acid or phosphatidylinositol bisphosphate. Co-immunoprecipitation indicates a close PLD2-Fes physical interaction that is negated by a Fes-R483K mutant that incapacitates its Src homology 2 domain. All these suggest for the first time the following mechanism: mTOR/S6K down-regulation→PLD2 overexpression→PLD2/Fes association→phosphatidic acid-led activation of Fes kinase→granulocytic differentiation. Differentiation shortening could have a clinical impact on reducing the time of return to normalcy of the white cell counts after chemotherapy in patients with acute promyelocytic leukemia.  相似文献   

18.
Both known isoforms of phospholipase (PL) D, PLD1 and PLD2, require phosphatidylinositol 4,5-bisphosphate for activity. However, PLD2 is fully active in the presence of this phospholipid, whereas PLD1 activation is dependent on additional factors such as ADP-ribosylation factor-1 (ARF-1) and protein kinase Calpha. We find that mastoparan, an activator of G(i) and mast cells, stimulates an intrinsic PLD activity, most likely PLD2, in fractions enriched in plasma membranes from rat basophilic leukemia 2H3 mast cells. Overexpression of PLD2, but not of PLD1, results in a large increase in the mastoparan-inducible PLD activity in membrane fractions, particularly those enriched in plasma membranes. As in previous studies, expressed PLD2 is localized primarily in the plasma membrane and PLD1 in granule membranes. Studies with pertussis toxin and other agents indicate that mastoparan stimulates PLD2 independently of G(i), ARF-1, protein kinase C, and calcium. Kinetic studies indicate that mastoparan interacts synergistically with phosphatidylinositol 4,5-bisphosphate and that oleate, itself a weak stimulant of PLD2 at low concentrations, is a competitive inhibitor of mastoparan stimulation of PLD2. Therefore, mastoparan may be useful for investigating the regulation of PLD2, particularly in view of the well studied molecular interactions of mastoparan with certain other strategic signaling proteins.  相似文献   

19.
A protein complex (PC) composed of the MRP8 and MRP14 proteins has previously been shown to be a specific inhibitor of casein kinase I and II. This PC is expressed during the late stages of terminal differentiation induced in human promyelocytic HL-60 leukemia cells by 1 alpha,25-dihydroxyvitamin D3 and in human monocytic THP-1 leukemia cells by phorbol 12-myristate 13-acetate. This expression is associated with terminal cell differentiation because incubation of HL-60 cells with an agent or condition that causes suppression of growth but not induction of differentiation does not result in expression of the PC. At concentrations of 5-15 nM, the purified PC inhibited the growth of HL-60 cells and THP-1 cells, as well as other cell types belonging to different cell lineages. This growth inhibition was preceded by a reduction in [32P]phosphate incorporation and, at the higher PC concentrations, was associated with a reduction in [3H]thymidine, [3H]uridine, and [32S]methionine incorporation. The specific expression pattern and growth-inhibitory character of the PC suggests that the complex may have a role in suppressing cell growth during monomyelocytic terminal differentiation induced by specific chemical stimuli and during physiological and pathological events associated with monomyelocytic cell functions.  相似文献   

20.
Oxidative stress or signaling is widely implicated in apoptosis, ischemia and mitogenesis. Previously, our group reported that the hydrogen peroxide (H2O2)-dependent activation of phospholipase D2 (PLD2) in PC12 cells is involved in anti-apoptotic effect. However, the precise mechanism of PLD2 activation by H2O2 was not revealed. To find H2O2-dependent PLD2-regulating proteins, we immunoprecipitated PLD2 from PC12 cells and found that glyceraldehyde 3-phosphate dehydrogenase (GAPDH) coimmunoprecipitated with PLD2 upon H2O2 treatment. This interaction was found to be direct by in vitro reconstitution of purified GAPDH and PLD2. In vitro studies also indicated that PLD2-associated GAPDH was modified on its reactive cysteine residues. Koningic acid, an alkylator of GAPDH on catalytic cysteine residue, also increased interaction between the two proteins in vitro and enhanced PLD2 activity in PC12 cells. Blocking H2O2-dependent modification of GAPDH with 3-aminobenzamide resulted in the inhibition of the GAPDH/PLD2 interaction and attenuated H2O2-induced PLD2 activation in PC12 cells. From the results, we suggest that H2O2 modifies GAPDH on its catalytic cysteine residue not only to inactivate the dehydrogenase activity of GAPDH but also to endow GAPDH with the ability to bind PLD2 and the resulting association is involved in the regulation of PLD2 activity by H2O2.  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