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.
1. Although stimulation of mouse RAW 264.7 macrophages by UTP elicits a rapid increase in intracellular free Ca2+ ([Ca2+]i), phosphoinositide (PI) turnover, and arachidonic acid (AA) release, the causal relationship between these signalling pathways is still unclear. In the present study, we investigated the involvement of phosphoinositide-dependent phospholipase C (PI-PLC) activation, Ca2+ increase and protein kinase activation in UTP-induced AA release. The effects of stimulating RAW 264.7 cells with thapsigargin, which cannot activate the inositol phosphate (IP) cascade, but results in the release of sequestered Ca2+ and an influx of extracellular Ca2+, was compared with the effects of UTP stimulation to elucidate the multiple regulatory pathways for cPLA2 activation. 2. In RAW 264.7 cells UTP (100 microM) and thapsigargin (1 microM) caused 2 and 1.2 fold increases, respectively, in [3H]-AA release. The release of [3H]-AA following treatment with UTP and thapsigargin were non-additive, totally abolished in the Ca2+-free buffer, BAPTA (30 microM)-containing buffer or in the presence of the cPLA2 inhibitor MAFP (50 microM), and inhibited by pretreatment of cells with pertussis toxin (100 ng ml(-1)) or 4-bromophenacyl bromide (100 microM). By contrast, aristolochic acid (an inhibitor of sPLA2) had no effect on UTP and thapsigargin responses. 3. U73122 (10 microM) and neomycin (3 mM), inhibitors of PI-PLC, inhibited UTP-induced IP formation (88% and 83% inhibition, respectively) and AA release (76% and 58%, respectively), accompanied by a decrease in the [Ca2+]i rise. 4. Wortmannin attenuated the IP response of UTP in a concentration-dependent manner (over the range 10 nM-3 microM), and reduced the UTP-induced AA release in parallel. RHC 80267 (30 microM), a specific diacylglycerol lipase inhibitor, had no effect on UTP-induced AA release. 5. Short-term treatment with PMA (1 microM) inhibited the UTP-stimulated accumulation of IP and increase in [Ca2+]i, but had no effect on the release of AA. In contrast, the AA release caused by thapsigargin was increased by PMA. 6. The role of PKC in UTP- and thapsigargin-mediated AA release was shown by the blockade of these effects by staurosporine (1 microM), Ro 31-8220 (10 microM), Go 6976 (1 microM) and the down-regulation of PKC. 7. Following treatment of cells with SK&F 96365 (30 microM), thapsigargin-, but not UTP-, induced Ca2+ influx, and the accompanying AA release, were down-regulated. 8. Neither PD 98059 (100 microM), MEK a inhibitor, nor genistein (100 microM), a tyrosine kinase inhibitor, had any effect on the AA responses induced by UTP and thapsigargin. 9. We conclude that UTP-induced cPLA2 activity depends on the activation of PI-PLC and the sustained elevation of intracellular Ca2+, which is essential for the activation of cPLA2 by UTP and thapsigargin. The [Ca2+]i-dependent AA release that follows treatment with both stimuli was potentiated by the activity of protein kinase C (PKC). A pertussis toxin-sensitive pathway downstream of the increase in [Ca2+]i was also shown to be involved in AA release.  相似文献   

2.
We have investigated the role and mechanism of protein kinase C (PKC) isoforms in endothelin-1 (ET-1)-induced arachidonic acid (AA) release in cat iris sphincter smooth muscle (CISM) cells. ET-1 increased AA release in a concentration (EC50=8 nM) and time-dependent (t1/2=1.2 min) manner. Cytosolic phospholipase A2 (cPLA2), but not phospholipase C (PLC), is involved in the liberation of AA in the stimulated cells. This conclusion is supported by the findings that ET-1-induced AA release is inhibited by AACOCF3, quinacrine and manoalide, PLA2 inhibitors, but not by U-73122, a PLC inhibitor, or by RHC-80267, a diacylglycerol lipase inhibitor. A role for PKC in ET-1-induced AA release is supported by the findings that the phorbol ester, PDBu, increased AA release by 96%, that prolonged treatment of the cells with PDBu resulted in the selective down regulation of PKCalpha and the complete inhibition of ET-1-induced AA release, and that pretreatment of the cells with staurosporine or RO 31-8220, PKC inhibitors, blocked the ET-1-induced AA release. G?-6976, a compound that inhibits PKCalpha and beta specifically, blocked ET-1-induced AA release in a concentration-dependent manner with an IC50 value of 8 nM. Thymeatoxin (0.1 microM), a specific activator of PKCalpha, beta, and gamma induced a 150% increase in AA release. Treatment of the cells with ET-1 caused significant translocation of PKCalpha, but not PKCbeta, from cytosol to the particulate fraction. These results suggest that PKCalpha plays a critical role in ET-1-induced AA release in these cells. Immunochemical analysis revealed the presence of cPLA2, p42mapk and p44mapk in the CISM cells. The data presented are consistent with a role for PKCalpha, but not for p42/p44 mitogen-activated protein kinase (MAPK), in cPLA2 activation and AA release in ET-1-stimulated CISM cells since: (i) the PKC inhibitor, RO 31-8220, inhibited ET-1-induced AA release, cPLA2 phosphorylation and cPLA2 activity, but had no inhibitory effect on p42/p44 MAPK activation, (ii) genistein, a tyrosine kinase inhibitor, inhibited ET-1-stimulated MAPK activity but had no inhibitory effect on AA release in the ET-1-stimulated cells. We conclude that in CISM cells, ET-1 activates PKCalpha, which activates cPLA2, which liberates AA for prostaglandin synthesis.  相似文献   

3.
The signaling pathway for protein kinase C (PKC) activation and the role of PKC isoforms in LPS-induced nitric oxide (NO) release were studied in RAW 264.7 macrophages. The tyrosine kinase inhibitor genestein attenuated LPS-induced NO release and inducible nitric oxide synthase (iNOS) expression, as did the phosphoinositide-specific phospholipase C (PI-PLC) inhibitor U73122 and the phosphatidylcholine-specific phospholipase C (PC-PLC) inhibitor D609. LPS stimulated phosphatidylinositol (PI) hydrolysis and PKC activity in RAW cells; both were inhibited by genestein. The PKC inhibitors (staurosporine, calphostin C, Ro 31-8220, or Go 6976) or long-term 12-O-tetradecanoylphorbol 13-acetate (TPA) treatment also resulted in inhibition of LPS-induced NO release and iNOS expression. Western blot analysis showed expression of PKC-alpha, -betaI, -delta, -eta, and -zeta in RAW cells; down-regulation of PKC-alpha, -betaI, and -delta, but not -eta, was seen after long-term TPA treatment, indicating the possible involvement of one or all of PKC-alpha, -betaI, and -delta, but not -eta, in LPS-mediated effects. Treatment with antisense oligonucleotides for these isoforms further demonstrated the involvement of PKC-alpha, -betaI, and delta, but not -eta, in LPS responses. Stimulation of cells with LPS for 1 h caused activation of NF-kappaB in the nuclei by detection of NF-kappaB-specific DNA-protein binding; this was inhibited by genestein, U73122, D609, calphostin C, or antisense oligonucleotides for PKC-alpha, -betaI, and -delta, but not -eta. These data suggest that LPS activates PI-PLC and PC-PLC via an upstream tyrosine kinase to induce PKC activation, resulting in the stimulation of NF-kappaB DNA-protein binding, then initiated the expression of iNOS and NO release. PKC isoforms alpha, betaI, and delta were shown to be involved in the regulation of these LPS-induced events.  相似文献   

4.
We previously demonstrated that the anticancer agent and protein kinase C (PKC) inhibitor 7-hydroxystaurosporine (UCN-01) induces apoptosis independently of p53 and protein synthesis in HL60 cells. We now report the associated changes of PKC isoforms. PKCalpha, betaI, betaII, delta, and zeta activities were measured after immunoprecipitation of cytosols from UCN-01-treated HL60 cells. UCN-01 had no effect on PKCzeta and inhibited kinase activity of PKCbetaI, betaII, and delta. PKCalpha activity was initially inhibited at 1 h, and subsequently increased as cells underwent apoptosis 3 h after the beginning of UCN-01 treatment. Camptothecin (CPT) and etoposide (VP-16) also markedly enhanced PKCalpha activity during apoptosis in HL60 cells. However, CPT did not affect PKCbetaI, betaII and zeta, and activated PKCdelta. PKCalpha activation was not due to increased protein levels or proteolytic cleavage but was associated with PKCalpha autophosphorylation in vitro and increased phosphorylation in vivo. We also found that not only PKC delta but also PKC betaI was proteolytically activated in HL60 cells during apoptosis. The PKCalpha activation and hyperphosphorylation were abrogated by N-benzyloxycarbonyl-Val-Ala-Asp(O-methyl)-fluoromethylketone (z-VAD-fmk) under conditions that abrogated apoptosis. z-VAD-fmk also prevented PKCdelta and betaI proteolytic activation. Together these findings suggest that caspases regulate PKC activity during apoptosis in HL60 cells. At least two modes of activation were observed: hyperphosphorylation for PKCalpha and proteolytic activation for PKC delta and betaI.  相似文献   

5.
Protein kinase C (PKC) is a family of enzymes that are physiologically activated by 1,2-diacylglycerol (DAG) and other lipids. To date, 11 different isozymes, alpha, betaI, betaII, gamma, delta, epsilon, nu, lambda(iota), mu, theta and zeta, have been identified. On the basis of their structure and activators, they can be divided into three groups, two of which are activated by DAG or its surrogate, phorbol 12-myristate 13-acetate (PMA). PKC isozymes are remarkably different in number and prevalence in different cell lines and tissues. When activated, the isozymes bind to membrane phospholipids or to receptors that are located in and anchor the enzymes in a subcellular compartment. Some PKCs may also be activated in their soluble form. These enzymes phosphorylate serine and threonine residues on protein substrates, perhaps the best known of which are the myristoylated, alanine-rich C kinase substrate and nuclear lamins A, B and C. The enzymes clearly play a role in signal transduction, and, because of the importance of PMA as a tumor promoter, they are thought to affect some aspect of cell cycling. How PKC takes part in the regulation of cell transformation, growth, differentiation, ruffling, vesicle trafficking and gene expression, however, is largely unknown.  相似文献   

6.
Several G(i)-linked neurotransmitter receptors, including dopamine D2 receptors, act synergistically with Ca(2+)-mobilizing stimuli to potentiate release of arachidonic acid (AA) from membrane phospholipids. In brain, AA and its metabolites are thought to act as intracellular second messengers, suggesting that receptor-dependent potentiation of AA release may participate in neuronal transmembrane signaling. To study the molecular mechanisms underlying this modulatory response, we have now used Chinese hamster ovary cells transfected with rat D2-receptor cDNA, CHO(D2). Two antisense oligodeoxynucleotides corresponding to distinct cDNA sequences of cytosolic, AA-specific phospholipase A2 (cPLA2) were synthesized and added to cultures of CHO(D2) cells. Incubation with antisense oligodeoxynucleotides inhibited D2 receptor-dependent release of AA but had no effect on D2-receptor binding or D2 inhibition of cyclic AMP accumulation. In addition, pharmacological experiments showed that D2 receptor-dependent AA release was prevented by nonselective phospholipase inhibitors (such as mepacrine) but not by inhibitors of membrane-bound, non-AA-specific PLA2 (such as p-bromophenacyl bromide). cPLA2 is expressed in brain tissue. The results, showing that cPLA2 participates in receptor-dependent potentiation of AA release in CHO(D2) cells, suggest that this phospholipase may serve a similar signaling function in brain.  相似文献   

7.
Bacterial lipopolysaccharide (LPS) is a potent activator of cells of the macrophage/monocyte lineage. Two mature macrophage cell lines, P388D1 and RAW264.7, exhibit very different biological responses to LPS. Although RAW264.7 cells release arachidonic acid from phospholipid in response to LPS stimulation, P388D1 cells do not respond in this manner. However, LPS primes P388D1 cells to release arachidonic acid in response to other stimuli. The goal of this work is to contrast the biochemical events that occur in LPS-treated P388D1 and RAW264.7 macrophages. Enzyme assays indicate that LPS treatment induces the activation of cytosolic PLA2 in RAW264.7, but not in P388D1 cells. Phorbol ester (PMA), a receptor-independent stimulus, also fails to induce arachidonic acid release from P388D1 cells, suggesting that these cells may have a defect in the signal transduction machinery that is common to LPS and PMA. This hypothesis is supported by the observation that the expression of the LPS receptors CD14 and CD11b/CD18 is similar on P388D1 and RAW264.7 cells. Western blot analyses indicate that the erk kinases are activated upon LPS treatment of RAW264.7 but not P388D1 cells. LPS-induced arachidonic acid release is reduced in cells treated with the MEK inhibitor PD98059, suggesting that activated erk kinases mediate the phosphorylation and activation of cPLA2 in this system. Interestingly, the p42 isoform of erk (erk2) appears to be activated in resting P388D1 cells. This observation indicates that the MAP kinase cascade may be constitutively activated in P388D1 cells which may in turn limit their ability to respond to LPS. Together, these data provide evidence that mature macrophages from different sources can exhibit variable responses to LPS and highlight the danger of making generalizations regarding the effects of LPS on macrophages.  相似文献   

8.
The metabotropic glutamate receptor (mGluR) agonist 1-aminocyclopentane-1S,3R-dicarboxylic acid (ACPD) potentiated the accumulation of cyclic AMP induced by either beta-adrenergic receptor stimulation (isoproterenol) or direct activation of adenylyl cyclase (AC) with forskolin in rat cerebral cortical astrocytes grown in a defined medium. In contrast, ACPD inhibits the cyclic AMP response in astrocytes cultured in a serum-containing medium. Pharmacological characterization indicated that a group I mGluR, of which only mGluR5 is detectable in these cells, is involved in the potentiation of cyclic AMP accumulation. Potentiation was elicited by mGluR I agonists [e.g., (R,S)-3,5-dihydroxyphenylglycine (DHPG)], but not by mGluR II or III agonists; it was pertussis toxin resistant and abolished by procedures suppressing mGluR5 function (phorbol ester pretreatment or DHPG-induced receptor down-regulation). Nevertheless, it appears that products generated through the mGluR5 transduction pathway, such as elevated [Ca2+]i or activated protein kinase C (PKC), are not involved in the potentiation as it was not influenced by either the intracellular calcium chelator BAPTA-AM or the PKC inhibitor Ro 31-8220. An inhibitor of phospholipase C, U-73122, markedly attenuated mGluR5-activated phosphoinositide hydrolysis but did not significantly affect the DHPG potentiation of the cyclic AMP response. A mechanism is proposed in which the potentiating effect on AC could be mediated by free betagamma complex that is liberated after the agonist-bound mGluR5 interacts with its coupled G protein.  相似文献   

9.
Neutrophils from patients with Chronic Myeloid Leukemia (CML) exhibit defects in several functions. They also show altered phosphorylation-dephosphorylation patterns of several proteins on stimulation with phorbol myristate acetate--a direct activator of protein kinase C (PKC). Since PKC mediates several functions of the neutrophil, in this study we investigate the PKC isoform profile and subcellular distribution in normal and CML neutrophils in an attempt to elucidate their role in CML signalling. Our results show the presence of PKC alpha, betaI, betaII and delta in both the cell types. A distinct and clear signal was obtained for PKC alpha, the isoform reported to be absent or present in very low amounts in normal neutrophils. In addition, PKC alpha was present in significantly lower levels in CML neutrophils while the PKC delta isoform was found in significantly higher amounts in the CML cytosol as compared to that in normal cells. PKC alpha, betaI, betaII and delta isoforms could not be detected in the nucleus of unstimulated normal and CML neutrophils. The altered levels of PKC alpha and delta may be one of the causes for the defects in function exhibited by the leukemic cells.  相似文献   

10.
11.
Attachment of HeLa cells to gelatin induces the release of arachidonic acid (AA), which is essential for cell spreading. HeLa cells spreading in the presence of extracellular Ca2+ released more AA and formed more distinctive lamellipodia and filopodia than cells spreading in the absence of Ca2+. Addition of exogenous AA to cells spreading in the absence of extracellular Ca2+ restored the formation of lamellipodia and filopodia. To investigate the role of cytosolic phospholipase A2 (cPLA2) in regulating the differential release of AA and subsequent formation of lamellipodia and filopodia during HeLa cell adhesion, cPLA2 phosphorylation and translocation from the cytosol to the membrane were evaluated. During HeLa cell attachment and spreading in the presence of Ca2+, all cPLA2 became phosphorylated within 2 min, which is the earliest time cell attachment could be measured. In the absence of extracellular Ca2+, the time for complete cPLA2 phosphorylation was lengthened to <4 min. Maximal translocation of cPLA2 from cytosol to membrane during adhesion of cells to gelatin was similar in the presence or absence of extracellular Ca2+ and remained membrane associated throughout the duration of cell spreading. The amount of total cellular cPLA2 translocated to the membrane in the presence of extracellular Ca2+ went from <20% for unspread cells to >95% for spread cells. In the absence of Ca2+ only 55-65% of the total cPLA2 was translocated to the membrane during cell spreading. The decrease in the amount translocated could account for the comparable decrease in the amount of AA released by cells during spreading without extracellular Ca2+. Although translocation of cPLA2 from cytosol to membrane was Ca2+ dependent, phosphorylation of cPLA2 was attachment dependent and could occur both on the membrane and in the cytosol. To elucidate potential activators of cPLA2, the extracellular signal-related protein kinase 2 (ERK2) and protein kinase C (PKC) were investigated. ERK2 underwent a rapid phosphorylation upon early attachment followed by a dephosphorylation. Both rates were enhanced during cell spreading in the presence of extracellular Ca2+. Treatment of cells with the ERK kinase inhibitor PD98059 completely inhibited the attachment-dependent ERK2 phosphorylation but did not inhibit cell spreading, cPLA2 phosphorylation, translocation, or AA release. Activation of PKC by phorbol ester (12-O-tetradecanoylphorbol-13-acetate) induced and attachment-dependent phosphorylation of both cPLA2 and ERK2 in suspension cells. However, in cells treated with the PKC inhibitor Calphostin C before attachment, ERK2 phosphorylation was inhibited, whereas cPLA2 translocation and phosphorylation remained unaffected. In conclusion, although cPLA2-mediated release of AA during HeLa cell attachment to a gelatin substrate was essential for cell spreading, neither ERK2 nor PKC appeared to be responsible for the attachment-induced cPLA2 phosphorylation and the release of AA.  相似文献   

12.
We examined the ability of opsonized zymosan (OPZ) to stimulate translocation of protein kinase C (PKC) isoforms in human neutrophils. Neutrophils express five PKC isoforms (alpha, betaI, betaII, delta, and zeta), but little is known of their individual roles in neutrophil activation. As determined by immunoblotting, OPZ caused a time-dependent translocation of the predominant PKC isoforms (betaII, delta, and zeta) to neutrophil membranes, with a concomitant loss from the cytosol. Maximal translocation of all three isoforms occurred by 3 min. No PKC immunoreactivity was observed in a crude nuclear fraction, but PKC-delta and -zeta were found in the granule fraction after degranulation (10 min). PKC activity (Ca2+-dependent and -independent) increased 50- and 19-fold, respectively, by 10 min in the granules from OPZ-stimulated cells. Curiously, no immunoreactive cPKC (alpha and beta(I/II)) could be localized in the granule fraction to account for the Ca2+-dependent PKC activity. Localization of PKC isoforms in the neutrophil membranes and granules suggests their involvement in the regulation of functional responses triggered by OPZ. PKC isoform translocation to membranes from OPZ-stimulated cells preceded both p47phox (a cytosolic component of the NADPH oxidase) translocation and NADPH oxidase assembly. The presence of both PKC isoforms and p47phox in the membrane was transient, with the loss of p47phox occurring sooner than either the loss of membrane-associated PKC or that of NADPH oxidase activity. The apparent EC50 values for PKC translocation and NADPH oxidase assembly were similar. These data suggest that PKC isoforms regulate the assembly and activation of NADPH oxidase induced by OPZ.  相似文献   

13.
This study was performed to examine effects of the overexpression of protein kinase C (PKC) isoforms (i.e., beta I, beta II, gamma, delta, eta, and zeta) on mitogen-activated protein (MAP) kinase (Erk-1 and -2) signaling and growth characteristics of NIH3T3 cells. Phorbol ester (PMA) activated endogenous and ectopically expressed PKC alpha, beta I, beta II, gamma, delta, epsilon, and eta. Overexpression of the examined PKC isoforms enhanced PMA-induced MAP kinase activation. Potentiation of MAP kinase activation was also observed upon stimulation of cells with platelet-derived growth factor (PDGF) although there was no indication for the activation PKC isoforms by PDGF. Inhibition of PKC blocked PMA- but not PDGF-induced MAP kinase activation. Thus, potentiation of PDGF-induced MAP kinase activation appears to be independent to PKC activity, while PMA-induced MAP kinase activation requires PKC activity. The ability of PKC isoforms to potentiate MAP kinase activation is not related to the growth characteristics of cells because individual PKC isoforms differentially regulated maximum density and proliferation of cells.  相似文献   

14.
The expression of protein kinase C (PKC) isozymes in human basophils and the regulation of PKC isozymes during basophil activation by phorbol 12-myristate 13-acetate (PMA) +/- ionomycin, f-met-leu-phe (FMLP), and anti-IgE antibody were examined. In human basophils (> 98% purity), PKCbetaI, betaII, delta, and were expressed, PKCalpha was difficult to detect, and PKCgamma and eta were undetectable. In unstimulated basophils, PKCbetaI and betaII were found primarily in the cytosol fraction (95% +/- 3% of total and 98% +/- 1%, respectively). Within 5 minutes of stimulation with PMA (100 ng/mL), both PKCbetaI and betaII were translocated to the membrane fraction (85% +/- 4% and 83% +/- 6%, respectively). In resting cells, 48% +/- 3% and 61% +/- 10% of PKCdelta and , respectively, existed in the membrane fraction. Within 1 minute of stimulation with PMA, 90% +/- 6% of PKC was found in the membrane fraction, however, no translocation of PKCdelta was apparent. Stimulation with FMLP caused modest translocation ( approximately 20%) of all PKC isozymes by 1 minute, whereas stimulation with anti-IgE antibody led to no detectable changes in PKC location throughout a 15-minute period of measurement. However, concentrations of PMA and ionomycin that alone caused no PKC translocation and little histamine release, together caused significant histamine release but no apparent PKC translocation. Studies with bis-indolylmaleimide analogs showed inhibition of PMA-induced, but not anti-IgE-induced, histamine release. These pharmacological studies suggest that PKC does not play a prodegranulatory role in human basophil IgE-mediated secretion.  相似文献   

15.
Protein kinase C (PKC) family members play significant roles in a variety of intracellular signal transduction processes, but information about the substrate specificities of each PKC family member is quite limited. In this study, we have determined the optimal peptide substrate sequence for each of nine human PKC isozymes (alpha, betaI, betaII, gamma, delta, epsilon, eta, mu, and zeta) by using an oriented peptide library. All PKC isozymes preferentially phosphorylated peptides with hydrophobic amino acids at position +1 carboxyl-terminal of the phosphorylated Ser and basic residues at position -3. All isozymes, except PKC mu, selected peptides with basic amino acids at positions -6, -4, and -2. PKC alpha, -betaI, -betaII, -gamma, and -eta selected peptides with basic amino acid at positions +2, +3, and +4, but PKC delta, -epsilon, -zeta, and -mu preferred peptides with hydrophobic amino acid at these positions. At position -5, the selectivity was quite different among the various isozymes; PKC alpha, -gamma, and -delta selected peptides with Arg at this position while other PKC isozymes selected hydrophobic amino acids such as Phe, Leu, or Val. Interestingly, PKC mu showed extreme selectivity for peptides with Leu at this position. The predicted optimal sequences from position -3 to +2 for PKC alpha, -betaI, -betaII, -gamma, -delta, and -eta were very similar to the endogenous pseudosubstrate sequences of these PKC isozymes, indicating that these core regions may be important to the binding of corresponding substrate peptides. Synthetic peptides based on the predicted optimal sequences for PKC alpha, -betaI,-delta, -zeta, and -mu were prepared and used for the determination of Km and Vmax for these isozymes. As judged by Vmax/Km values, these peptides were in general better substrates of the corresponding isozymes than those of the other PKC isozymes, supporting the idea that individual PKC isozymes have distinct optimal substrates. The structural basis for the selectivity of PKC isozymes is discussed based on residues predicted to form the catalytic cleft.  相似文献   

16.
Megakaryocytes undergo a unique differentiation program, becoming polyploid through repeated cycles of DNA synthesis without concomitant cell division. We have shown previously that phorbol 12-myristate 13-acetate (PMA) induces the Dami human megakaryocytic cell line to become polyploid and to express platelet-specific proteins, including von Willebrand factor (vWF) and glycoprotein Ib (GpIb). Phorbol esters are thought to regulate gene expression principally through the activation of protein kinase C (PKC), a family of structurally related kinases with potentially unique activation requirements and substrate specificities. A survey of PKC isoforms in Dami cells revealed that, by both Western and Northern analyses, PKC isoforms alpha, beta, delta, epsilon, eta, theta, and zeta were reproducibly detected. PKC-gamma was not detected. In order to define the role of individual PKC isoforms in megakaryocytic maturation, PMA and 2-deoxyphorbol 13-phenylacetate 20-acetate (dPPA), a putative selective activator of the PKC-beta 1 isotype, were compared for their effects on Dami cell maturation. Treatment with either dPPA or PMA caused Dami cells to cease proliferating, to become polyploid, and to express vWF. We also examined dPPA and PMA for their ability to activate and to downregulate expression of different PKC isoforms. Fifteen-minute treatment with PMA resulted in the translocation of PKC isoforms alpha, epsilon, and theta from the cytosolic to the membrane fraction; twenty-four hour treatment resulted in the downregulation of these isoforms. In contrast, dPPA was found to be a potent activator of PKC-epsilon alone and exhibited weaker effects on alpha and theta. These data suggest that PKC isoforms beta, delta, eta, and zeta, which appear not to be activated by either phorbol ester, are unlikely to be primarily involved in megakaryocytic maturation in response to these agents. The isoforms that are translocated by both phorbol esters-PKC isoforms alpha and theta, and particularly epsilon-are more likely to transduce the signals that stimulate Dami cell differentiation.  相似文献   

17.
BACKGROUND: FRTL-5 thyroid cells are a cell line extensively used for the investigation of thyroid functions. Activation of alpha-1 adrenergic receptors stimulates both arachidonic acid (AA) release and cytosolic Ca2+ increase in this cell line. Cytosolic Ca2+ and arachidonic acid are known to be important second messengers regulating a variety of thyroid functions. The generation of these messengers is regulated primarily by two different types of phospholipases, phospholipase C (PLC) and phospholipase A2 (PLA2). METHODS: Norepinephrine (NE, 10 mumol/L) was used as an alpha-1 adrenergic activator, and cytosolic-free Ca2+ concentration ([Ca2+]i) was determined using the fluorescent dye indo-1. Arachidonic acid release was measured as an indicator of PLA2 activation, and protein kinase C (PKC) activity determination and isoforms identification were performed using commercial kits. RESULTS: Norepinephrine increased [Ca2+]i and AA release. Prevention of NE-induced cytosolic Ca2+ influx, either by removal of extracellular Ca2+ or by use of Ca2+ channel blockers, NiCl2 or CoCl2, inhibited AA generation entirely. Inhibition of NE-induced increase in [Ca2+]i by the Ca2+ chelator, 1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid (BAPTA), also significantly suppressed NE-induced AA release. Inhibition of PKC activity by PKC inhibitors (H-7 or staurosporine) or downregulation induced by prolonged treatment with phorbol 12-myristate 13-acetate (PMA) or thyleametoxin (TX) significantly blocked the NE-induced AA release, which indicates PKC is involved in mediating NE-induced AA release. Protein kinase C activity measurement indicated that NE induced an activation of PKC in 5 minutes. To further characterize the role of PKC or Ca2+ in regulation of AA release, we identified PKC isoforms by immunoblotting with specific antibodies against 8 different Protein kinase C isoforms. PKC-alpha, -beta I, -beta II, -gamma, delta, -epsilon, -zeta, and -eta isoforms were identified. Norepinephrine induced translocation of PKC-alpha, -beta I, -beta II, -gamma, -delta, and -epsilon isoforms but not -zeta and -eta from cytosol to membrane. Chelation of intracellular Ca2+, prevention of Ca2+ influx, or prolonged treatment with thymeleatoxin (TX) completely blocked the NE-induced translocation of PKC-alpha. CONCLUSIONS: These results, taken together with data obtained from AA experiments, suggest that PKC plays a critical role in alpha-1 adrenergic receptor mediated PLA2 activation and subsequent AA release. Extracellular Ca2+ influx is a prerequisite for both PKC-alpha translocation and AA release. Whether Ca2+ acts directly upon the PLA2, or via PKC-alpha, to regulate AA generation is an intriguing question that remains to be clarified.  相似文献   

18.
Exposure of the gonadotrope cells to gonadotropin-releasing hormone (GnRH) reduces their responsiveness to a new GnRH stimulation (homologous desensitization). The time frame as well as the mechanisms underlying this phenomenon are yet unclear. We studied in a gonadotrope cell line (alphaT3-1) the effects of short as well as long term GnRH pretreatments on the GnRH-induced phospholipases-C (PLC), -A2 (PLA2) and -D (PLD) activities, by measuring the production of IP3, total inositol phosphates (IPs), arachidonic acid (AA) and phosphatidylethanol (PEt) respectively. We demonstrated that although rapid desensitization of GnRH-induced IP3 formation did not occur in these cells, persistent stimulation of cells with GnRH or its analogue resulted in a time-dependent attenuation of GnRH-elicited IPs formation. GnRH-induced IPs desensitization was potentiated after direct activation of PKC by the phorbol ester TPA, suggesting the involvement of distinct mechanisms in the uncoupling exerted by either GnRH or TPA on GnRH-stimulated PI hydrolysis. The levels of individual phosphoinositides remained unchanged under any desensitization condition applied. Interestingly, while the GnRH-induced PLA2 activity was rapidly desensitized (2.5 min) after GnRH pretreatments, the neuropeptide-evoked PLD activation was affected at later times, indicating an important time-dependent contribution of these enzymatic activities in the sequential events underlying the GnRH-induced homologous desensitization processes in the gonadotropes. Under GnRH desensitization conditions, TPA was still able to induce PLD activation and to further potentiate the GnRH-evoked PLD activity. AlphaT3-1 cells possess several PKC isoforms which, except PKCzeta, were differentially down-regulated by TPA (PKCalpha, betaII, delta, epsilon, eta) or GnRH (PKCbetaII, delta, epsilon, eta). In spite of the presence of PKC inhibitors or down-regulation of PKC isoforms by TPA, the desensitizing effect of the neuropeptide on GnRH-induced IPs, AA and PEt formation remained unchanged. In conclusion, in alphaT3-1 cells the GnRH-induced homologous desensitization affects the GnRH coupling with PLC, PLA2 and PLD by mechanism(s) which do not implicate TPA-sensitive PKC isoforms, but likely reflect time-dependent modification(s) on the activation processes of the enzymes.  相似文献   

19.
The goal of this study was to explain the priming effect of lipopolysaccharides (LPS) in human polymorphonuclear leukocytes on leukotriene B4 (LTB4) biosynthesis after stimulation with the receptor-mediated agonist formyl-methionyl-leucyl-phenylalanine (fMLP). This priming effect for LTB4 biosynthesis was maximal after a 30 min preincubation with LPS but was lost when incubations were extended to 90 min or longer. Priming with LPS resulted in an enhanced maximal activation of 5-lipoxygenase (5- to15-fold above unprimed cells) as well as a prolonged activation of the enzyme after stimulation with fMLP compared to that measured in unprimed cells. The activation of 5-lipoxygenase was associated with its translocation to the nuclear fraction of the cell after stimulation of LPS-primed cells but not of unprimed cells. Priming of cells with LPS also resulted in an enhanced capacity (fivefold increase) for arachidonic acid (AA) release after stimulation with fMLP compared to unprimed cells as measured by mass spectrometry. This release of AA was very efficiently blocked in a dose-dependent manner by the 85 kDa cytosolic phospholipase A2 (PLA2) inhibitor MAFP (IC50=10nM) but not by the 14 kDa secretory PLA2 inhibitor SB 203347 (up to 5 microM), indicating that the 85 kDa cPLA2 is the PLA2 responsible for AA release in response to receptor-mediated agonists. In accord with inhibitor studies, the LPS-mediated phosphorylation of cPLA2 followed the same kinetics as the priming for AA release, and a measurable fMLP-induced translocation of cPLA2 was observed only in primed cells. As with AA release and LTB4 biosynthesis, both the phosphorylation and capacity to translocate cPLA2 were reversed when the preincubation period with LPS was extended to 120 min. These results explain some of the cellular events responsible for the potentiation and subsequent decline of functional responses of human polymorphonuclear leukocytes recruited to inflammatory foci.  相似文献   

20.
Prostaglandins are known to play a central role in the initiation of labor in humans, and amnionic cells constitute a major source of these compounds. Prostaglandin synthesis and release by amnion cells in response to hormones and ligands takes place after a characteristic 4-5 h lag. However, we report herein that free arachidonic acid (AA), the metabolic precursor of prostaglandins, can be induced at much shorter times (1 h) in human amnionic WISH cells by phorbol 12-myristate 13-acetate (PMA) through activation of protein kinase Calpha (PKCalpha). WISH cells were found to possess both cytosolic group IV phospholipase A2 (cPLA2) and Group VI Ca2+-independent phospholipase A2 (iPLA2). Of these, the cPLA2 was found to be the likely mediator of AA mobilization in PMA-activated WISH cells. PMA also activates phospholipase D (PLD) in these cells and ethanol, a compound that inhibits PLD-mediated phosphatidic acid (PA) formation, blocked AA release. Moreover, prevention of PA dephosphorylation by the PA phosphohydrolase inhibitors propranolol and bromoenol lactone, resulted in inhibition of AA release by PMA-treated WISH cells. Collectively, these data suggest that activation of cPLA2 and attendant AA release by phorbol esters in WISH cells requires prior generation of DAG by phosphatidate phosphohydrolase.  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