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 406 毫秒
1.
M Ohmichi  S J Decker  L Pang  A R Saltiel 《Biochemistry》1992,31(16):4034-4039
The protein kinase inhibitors staurosporine and K252A inhibit some of the cellular actions of nerve growth factor (NGF). To explore the molecular mechanisms involved, we test the ability of these agents to block one of the earliest cellular responses to NGF, protein tyrosine phosphorylation. Concentrations of 10-100 nM staurosporine and K252A inhibit NGF-dependent tyrosine phosphorylation in PC12 cells and inhibit trk oncogene-dependent tyrosine phosphorylation in trk-transformed NIH3T3 (trk-3T3 cells). In contrast, these compounds are without effect on epidermal growth factor (EGF)-stimulated tyrosine phosphorylation in PC12 cells. NGF-stimulated tyrosine phosphorylation of the pp140c-trk NGF receptor and tyrosine phosphorylation of pp70trk are also inhibited by similar concentrations of staurosporine and K252A, whereas tyrosine phosphorylation of the EGF receptor, insulin receptor, and v-src is not affected. Both staurosporine and K252A inhibit the autophosphorylation of pp70trk on tyrosine residues in an in vitro immune complex kinase reaction. Incubation of trk-3T3 cells with 10 nM staurosporine causes rounded transformed cells to revert to a normal flattened phenotype, whereas src-transformed cells are unaffected by this agent. These data suggest that staurosporine and K252A specifically inhibit the trk tyrosine kinase activity through a direct mechanism, probably accounting for the attenuation by these agents of the cellular actions of NGF.  相似文献   

2.
The protein kinase C (PKC) inhibitor staurosporine was found to dramatically alter the actin microfilament cytoskeleton of a variety of cultured cells, including PTK2 epithelial cells, Swiss 3T3 fibroblasts, and human foreskin fibroblasts. For example, PTK2 cells exposed to 20 nM staurosporine exhibited a progressive thinning and loss of cytoplasmic actin microfilament bundles over a 60-min period. During this time microtubule and intermediate filament systems remained intact (as shown by immunofluorescence and at higher resolution by photoelectron microscopy), and the cells remained spread even though microfilament bundles were absent. Higher doses of staurosporine or longer exposure times at lower doses resulted in morphological alterations, but even severely arborized cells recovered normal morphology and actin patterns after a wash and an incubation for several hours in fresh medium. The actin filament disruption induced by staurosporine was distinguishable from the actin reorganization induced by exposure to the tumor promoter (and activator of PKC) phorbol myristate acetate (PMA). Swiss 3T3 cells made deficient in PKC by prolonged exposure to PMA (PKC down-regulation) exhibited actin alterations in response to staurosporine which were comparable to those in cells which had not been exposed to the phorbol ester. In a parallel control experiment, the actin cytoskeleton of PKC-deficient 3T3 cells was unaffected in response to PMA, consistent with down-regulation of this kinase. While the exact mechanism of staurosporine-induced actin reorganization remains to be determined, the observed effects of staurosporine on PKC-deficient cells make a role for PKC unlikely. These results indicate the need for care when staurosporine is employed as an inhibitor of protein kinase C in studies involving intact cells.  相似文献   

3.
In rabbit peritoneal neutrophils prelabeled with [3H] lyso platelet-activating factor, a protein kinase C inhibitor, staurosporine (> 1 microM), increased [3H]phosphatidylethanol ([3H]PEt) level in the presence of ethanol in a concentration- and time-dependent manner, providing evidence for staurosporine activation of phospholipase D (PLD). The staurosporine activation of the enzyme absolutely required both extracellular calcium and cytochalasin B, and was almost completely inhibited by pretreatment of the cells with pertussis toxin (IAP). In a reconstituted system where the purified Gi1 had been incorporated into phospholipid vesicles, staurosporine activated GTPase activity of Gi1 in a concentration-dependent fashion, with a maximal 4-5-fold effect. ADP-ribosylation by IAP of Gi1 in vesicles significantly suppressed the staurosporine activation. As with the GTPase activity of Gi1, GTPase activities of other purified IAP-sensitive G proteins, such as Gi2 and G(o), were significantly stimulated by staurosporine, but the cholera toxin substrate Gs was appreciably less sensitive to the staurosporine stimulation. The staurosporine activation of GTPase was also observed in rabbit neutrophil membranes from control cells, but not in membranes from IAP-treated neutrophils. From these results, we conclude that the staurosporine activation of PLD in rabbit neutrophils is attributed to the direct activation of an IAP-sensitive G protein in a similar manner to receptors occupied by agonists. By contrast, staurosporine failed to activate phosphoinositide-specific phospholipase C (PI-PLC) under the conditions in which it activated PLD, indicating that there exists a PLD activation pathway independent of PI-PLC. Furthermore, it was found that N-acetyl-beta-glucosaminidase release from the granules of intact neutrophils was evoked by staurosporine to almost the same extent as by fMLP (100 nM), but O2- generation was not affected. These results suggest a possibility that PLD pathway plays an important role in enzyme release, but is not sufficient for O2- generation, in rabbit peritoneal neutrophils.  相似文献   

4.
Abstract: We characterized the activation of interleukin-1β-converting enzyme (ICE)-like proteases (caspases) in human neuroblastoma cells (SH-SY5Y) following challenge with staurosporine, an established agent known to induce apoptosis. Time course analyses of lactate dehydrogenase release detected a significant increase in cell death as early as 6 h that continued at least until 24 h following staurosporine treatment. Western blot analyses using anti-poly(ADP-ribose) polymerase (anti-PARP) and anti-CPP32 antibodies revealed proteolytic processing of CPP32 (an ICE homologue) as well as fragmentation of PARP as early as 3 h following staurosporine challenge. Furthermore, the hydrolysis of the CPP32 substrate acetyl-DEVD-7-amido-4-methylcoumarin was detected as early as 3 h and became maximal at 6 h after staurosporine challenge, suggesting a delayed and sustained period of CPP32-like activation. In addition, we used the first immunohistochemical examination of CPP32 and PARP in cells following an apoptotic challenge. The localization of CPP32 in untreated SH-SY5Y cells was exclusively restricted to the cytoplasm. Following staurosporine challenge there was a condensing of CPP32 immunofluorescence from the cytoplasm to a region adjacent to the plasma membrane. In contrast, PARP immunofluorescence was evenly distributed in the nucleus in untreated SH-SY5Y cells and on staurosporine challenge was found to be associated with condensed chromatin. It is important that a pan ICE inhibitor [carbobenzoxy-Asp-CH2OC(O)-2,6-dichlorobenzene] was able to attenuate lactate dehydrogenase release and PARP and CPP32 cleavage and altered immunohistochemical staining patterns for PARP and CPP32 following staurosporine challenge.  相似文献   

5.
Apoptosis of cholinergic neurons is one of the core hallmarks of Alzheimer’s disease. SH-SY5Y neuroblastoma cells differentiated to the cholinergic phenotype were exposed to 100 nM staurosporine. Over a treatment period of 24 h, the pro- and anti-apoptotic factors, caspase-3 and Bcl-2, as well as LDH release as a measure of cell viability, were assessed in conjunction with the number of apoptotic cells by means of fluorescence-activated cell sorting. Caspase-3 activity and LDH release increased by 30% and 20% over controls, respectively, while Bcl-2 levels rose by 200% over controls. Furthermore, staurosporine treatment resulted in decreased acetylcholinesterase (AChE) enzymatic activity and decreased protein levels of the AChE splice variant tailed AChE (AChE-T). Only a slight increase in levels of readthrough AChE (AChE-R) was observed. Likewise, staurosporine reduced levels and activity of the cholinergic players choline acetyltransferase and high affinity choline uptake. The present study demonstrates that treatment with staurosporine leads to apoptotic events, which, however, are not reflected in the increased AChE activity and the alterations of AChE isoforms expression that are usually seen in apoptotic conditions. The effects of various additional phosphorylation inhibitors on AChE activity suggest that these unexpected cholinergic effects, firstly, are linked to the impact of staurosporine on phosphorylation and, secondly, reveal themselves in a first phase of cellular adaption that precedes neurotoxicity and subsequent cell death.  相似文献   

6.
The staurosporine-induced G1 cell cycle arrest was analyzed in a variety of cell lines which includes human tumor cell lines and oncogene-transformed NIH3T3 cell lines. All the cell lines which were sensitive to staurosporine-induced G1 arrest contained a functional retinoblastoma protein (pRB). However, when pRB-lacking fibroblast cells derived from pRB knockout mice were tested they were also sensitive to G1 arrest by staurosporine, indicating that the inactivation of pRB alone is not sufficient for the abrogation of staurosporine-induced G1 arrest. In searching for a common event caused by staurosporine, the cyclin-dependent kinase (CDK) inhibitor protein p27kip1but not p21cip1was found to accumulate after staurosporine treatment in all the cell lines examined. This accumulation occurred regardless of the induction of the G1 arrest. The result indicates that the accumulation of p27kip1is the cell's primary response to staurosporine and that the capability of staurosporine to induce G1 arrest depends on the integrity of cell cycle regulatory components which are downstream of p27kip1.  相似文献   

7.
Apoptosis was induced in embryonic chick cardiomyocytes by staurosporine. Treatment of cardiomyocytes with the preferential caspase-2 inhibitor, z-VDVAD-fmk (100 microM), produced a significant (P < 0.05) although small reduction in the amount of cell death. Ac-DVED-cmk (100 microM), which preferentially inhibits caspase-3 but inhibits to a lesser extent caspase-6, -7, -8, and -10, produced a minimal decrease in cell death. The combination of the caspase-3 and -2 inhibitors produced an additive reduction in cell death after staurosporine (1 microM for 6 h) from 80.4 +/- 0.7 to 54.6 +/- 1.3%. The ability of staurosporine to activate caspase-3 was confirmed in these cardiomyocytes by measurement of caspase-3 activity. A role for ceramide formation, from sphingomyelin to induce caspase activation was unlikely, as there were no changes in cellular ceramide or sphingomyelin after staurosporine treatment of cardiomyocytes when sphingomyelin was labeled by [(3)H]palmitate for 24 h. Neither were there any changes in sphingomyelinase activity. While staurosporine effectively suppressed PKC activity, phorbol 12-myristate 13 acetate did not alter staurosporine-induced cell death or DNA fragmentation. These results demonstrate that, in this model of cardiac cell death, caspase-2 inhibition is of considerable importance, caspase-3 inhibition is of lesser significance but may produce additional effects in the combination with caspase-2 inhibition, and ceramide production from sphingomyelin is not operative in the pathway leading to caspase activation and cell death.  相似文献   

8.
Over a decade ago, the gene STT3 was identified in a staurosporine and temperature sensitivity screen of yeast. Subsequently the product of this gene was shown to be a subunit of the endoplasmic reticulum-localized oligosaccharyl transferase (OT) complex. Although stt3 mutants are known to be staurosporine-sensitive, we found that mutants of other OT subunits (except ost4 Delta) are staurosporine-resistant, which indicates that this phenotype of stt3 mutants is not simply a consequence of their defect in glycosylation, as previously speculated. Staurosporine sensitivity was found to be an allele-specific phenotype restricted to cells harboring mutations in highly conserved residues in the N-terminal domain of the STT3 protein. Cells bearing mutations in one of the cytosolic-oriented loops (amino acids 158-168) in the N terminus of Stt3p were found to be specifically susceptible to staurosporine. Staurosporine is a specific inhibitor of Pkc1p, and a genetic link had previously been suggested between PKC1 and STT3. It is known that overexpression of PKC1 suppresses the staurosporine sensitivity of the stt3 mutants in an allele-specific manner, which is typical of mutants of Pkc1p cascade. It has been shown that the pkc1 null mutant exhibits lowered OT activity. Our results combined with these previous observations indicate that the N-terminal domain of Stt3p may interact with members of the Pkc1p cascade and consequently mutations in this domain result in staurosporine sensitivity. We further speculate that the Pkc1p regulates OT activity through the N-terminal domain of Stt3p, the C-terminal domain of which possesses the recognition and/or catalytic site of the OT complex.  相似文献   

9.
Ligand stimulation of the platelet-derived growth factor receptor (PDGF-R) results in rapid activation of the receptor tyrosine kinase, stimulation of phosphoinositide hydrolysis, an increase in intracellular free Ca2+ concentration ([Ca2+]i), and, ultimately, cellular proliferation. In a previous study, we demonstrated that staurosporine, a known inhibitor of protein kinase C, blocked PDGF-induced [Ca2+]i increases in Swiss mouse 3T3 fibroblasts by a mechanism that appeared unrelated to inhibition of protein kinase activity (Olsen, R., Melder, D., Seewald, M., Abraham, R., and Powis, G. (1990) Biochem. Pharmacol. 39, 968-972). In the present study, we report that staurosporine inhibits ligand-dependent PDGF-R tyrosine kinase activation in cell-free receptor preparations and in intact Swiss 3T3 cells. At the same concentrations (10(-8)-10(-6) M), staurosporine suppressed both the tyrosine phosphorylation of phospholipase C activity and the hydrolysis of phosphoinositides induced by PDGF stimulation of intact cells. In contrast, guanine nucleotide-binding protein-dependent phospholipase C activation induced by bradykinin or fluoroaluminate anion was relatively insensitive to staurosporine. A preferential inhibitory effect of staurosporine on signal generation by the PDGF-R was indicated by findings that epidermal growth factor receptor (EGF-R) tyrosine kinase activity and EGF-dependent phospholipase C in A-431 carcinoma cells were approximately 100-fold less sensitive to this drug. These data indicate that submicromolar concentrations of staurosporine inhibit PDGF-dependent phosphoinositide hydrolysis and Ca2+ mobilization through a proximal inhibitory effect on ligand-induced activation of the PDGF-R tyrosine kinase.  相似文献   

10.
In this study, it was shown that the mammalian sterile 20-like serine/threonine protein kinase 3 (Mst3) plays an essential role in the staurosporine-induced apoptosis of HeLa cells. The staurosporine-induced apoptosis was reduced by around 65% by the selective knockdown of Mst3 in stable clones, HeLa(siMst3). Although caspases were shown to be involved in the Mst3-mediated apoptosis, only 15–20% of staurosporine-induced apoptosis was suppressed by the caspase inhibitor, z-DEVD-fmk. Accordingly, Mst3 was proposed to trigger a caspase-independent apoptotic pathway in response to staurosporine. Interestingly, staurosporine greatly induced the mitochondrial membrane potential transition in HeLa cells, but had no effect in Hela(siMst3). The role of Mst3 in controlling the mitochondrial integrity was therefore proposed, presumably through the regulation of Bax. Furthermore, it was shown that staurosporine promoted the nuclear translocation of apoptosis-inducing factor and endonuclease G in HeLa cells. The nuclease activity associated with endonuclease G was also enhanced in response to staurosporine. However, both staurosporine-induced nuclear translocation of apoptosis-inducing factor and endonuclease G and the nuclease activity associated with endonuclease G were markedly reduced in Hela(siMst3). These results suggest that Mst3 may respond to staurosporine to trigger the caspase-independent apoptotic pathway by regulating the nuclear translocation of apoptosis-inducing factor and endonuclease G, and the nuclease activity associated with endonuclease G.  相似文献   

11.
To test the role of caspase 3 in apoptosis and in overall cell lethality caused by the protein kinase inhibitor staurosporine, we compared the responses of MCF-7c3 cells that express a stably transfected CASP-3 gene to parental MCF-7:WS8 cells transfected with vector alone and lacking procaspase-3 (MCF-7v). Cells were exposed to increasing doses (0.15-1 microM) of staurosporine for periods up to 19 h. Apoptosis was efficiently induced in MCF-7c3 cells, as demonstrated by cytochrome c release, processing of procaspase-3, procaspase-8, and Bid, increase in caspase-3-like DEVDase activity, cleavage of the enzyme poly(ADP-ribose) polymerase, DNA fragmentation, changes in nuclear morphology, and TUNEL assay and flow cytometry. For all of these measures except cytochrome c release, little or no activity was detected in MCF-7v cells, confirming that caspase-3 is essential for efficient induction of apoptosis by staurosporine, but not for mitochondrial steps that occur earlier in the pathway. MCF-7c3 cells were more sensitive to staurosporine than MCF-7v cells when assayed for loss of viability by reduction of a tetrazolium dye. However, the two cell lines were equally sensitive to killing by staurosporine when evaluated by a clonogenic assay. A similar distinction between apoptosis and loss of clonogenicity was observed for the cancer chemotherapeutic agent VP-16. These results support our previous conclusions with photodynamic therapy: (a) assessing overall reproductive death of cancer cells requires a proliferation-based assay, such as clonogenicity; and (b) the critical staurosporine-induced lethal event is independent of those mediated by caspase-3.  相似文献   

12.
Staurosporine is a microbial anti-fungal alkaloid having a most potent inhibitory activity on protein kinase C and is recently found as a non-12-O-tetradecanoylphorbol-13-acetate (non-TPA)-type tumor promoter of mouse skin, although tumor promotion induced by a TPA-type tumor promoter teleocidin is suppressed by staurosporine. When rat peritoneal macrophages were incubated in the medium containing various concentrations of staurosporine, prostaglandin E2 production and release of radioactivity from [3H]arachidonic acid-labeled macrophages were stimulated at concentrations of 1 and 10 ng/ml. But higher concentrations of staurosporine such as 100 and 1000 ng/ml showed no stimulative effect on prostaglandin E2 production although cytoplasmic free calcium levels were increased in a dose-dependent manner. Staurosporine-induced stimulation of prostaglandin E2 production was inhibited by treatment with cycloheximide, suggesting that a certain protein synthesis is prerequisite for the stimulation of arahcidonic acid metabolism. At higher concentrations (100 and 1000 ng/ml), staurosporine inhibited TPA-type tumor promoter (TPA, teleocidin and aplysiatoxin)-induced stimulation of arachidonic acid metabolism probably due to the inhibition of protein kinases. Tumor promotion activity and anti-tumor promotion activity of staurosporine might be explained by the fact that the lower concentrations of staurosporine stimulate arachidonic acid metabolism and the higher concentrations of staurosporine inhibit the tumor promoter-induced arachidonic acid metabolism, respectively.  相似文献   

13.
A digital imaging fluorescence microscope was used to study the effect of a protein kinase inhibitor staurosporine on the antigen-dependent calcium signals in an individual rat basophilic leukemia cell (RBL-2H3). Although dose dependency of staurosporine was different from one cell to another, staurosporine inhibited, at low concentration, the calcium influx from the external medium into RBL-2H3 cells. At high concentration, however, it inhibited both the removal of calcium ion from internal stores and the calcium influx from the external medium. These results indicated that staurosporine is necessary for the inhibition of the calcium influx from the external medium and that a protein kinase (possibly protein kinase C) is involved in the calcium influx from the external medium into the cytoplasm.  相似文献   

14.
Staurosporine, a microbial alkaloid, enhances inositol 1,4,5-trisphosphate (IP3) and 1,2-diacylglycerol (DG) production rapidly and dose-dependently in fMet-Leu-Phe (FMLP)-stimulated human neutrophils showing maximal effects at 1 microM concentration. The IP3 increase was specific for staurosporine as three other putative protein kinase C (PKC) inhibitors, H7, sphingosine and palmitoylcarnitine were unable to enhance the IP3 generation in FMLP-stimulated human neutrophils. Staurosporine, at concentrations 0.3-1.0 microM, did not affect the initial mobilization of FMLP-induced intracellular Ca2+ (Ca2+i), although a sustained elevation of cytosolic Ca2+ level was observed within 5 min. This effect could not be suppressed, even by 1 microM phorbol-myristate 12,13-acetate (PMA). Whereas lower concentrations of staurosporine (less than or equal to 100 nM) were unable to affect FMLP-induced IP3 production, DG accumulation and Ca2+i, the PMA-inhibited initial Ca2+i signal and IP3 formation triggered by FMLP were almost completely restored. At higher concentrations (greater than or equal to 300 nM) staurosporine reversed the inhibitory effect of other protein kinases, distinct from the PMA-inducible one, which may be responsible for the phosphatidyl inositol 4,5-bisphosphate (PIP2) breakdown, thus causing accumulation of IP3 and DG and an elevation of C2+i level. Whereas IP3 declined to basal level within 5 min, the DG level remained elevated during the same period. This phenomenon is attributed to phospholipase D (PLD) stimulation by staurosporine, which augments the DG synthesis, in part through PA degradation via phosphatidic acid (PA) phosphohydrolase.  相似文献   

15.
We show that mitochondrial DNA (mtDNA)-depleted 143B cells are hypersensitive to staurosporine-induced cell death as evidenced by a more pronounced DNA fragmentation, a stronger activation of caspase-3, an enhanced poly(ADP-ribose) polymerase-1 (PARP-1) cleavage, and a more dramatic cytosolic release of cytochrome c. We also show that B-cell CLL/lymphoma-2 (Bcl-2), B-cell lymphoma extra large (Bcl-X(L)), and myeloid cell leukemia-1 (Mcl-1) are constitutively less abundant in mtDNA-depleted cells, that the inhibition of Bcl-2 and Bcl-X(L) can sensitize the parental cell line to staurosporine-induced apoptosis, and that overexpression of Bcl-2 or Bcl-X(L) can prevent the activation of caspase-3 in ρ(0)143B cells treated with staurosporine. Moreover, the inactivation of cathepsin B with CA074-Me significantly reduced cytochrome c release, caspase-3 activation, PARP-1 cleavage, and DNA fragmentation in mtDNA-depleted cells, whereas the pan-caspase inhibitor failed to completely prevent PARP-1 cleavage and DNA fragmentation in these cells, suggesting that caspase-independent mechanisms are responsible for cell death even if caspases are activated. Finally, we show that cathepsin B is released in the cytosol of ρ(0) cells in response to staurosporine, suggesting that the absence of mitochondrial activity leads to a facilitated permeabilization of lysosomal membranes in response to staurosporine.  相似文献   

16.
Mature mouse oligodendrocytes (OLs) are susceptible to death in demyelinating diseases such as multiple sclerosis and in brain injury following neurotrauma, ischemia, or stroke. To understand mechanisms leading to death of mature OLs and develop strategies for protection, we utilized cultures of mature mouse OLs to investigate the role of caspases and calpains in OL cell death mediated by different mechanisms. The agents used were (i) staurosporine, which induces apoptotic death via inhibition of protein kinases; (ii) kainate, which activates non-NMDA glutamate receptors; (iii) thapsigargin, which releases intracellular calcium stores; and (iv) SNAP, which releases active NO species and causes necrotic cell death. Inhibitors blocking primary effector caspases (including caspase 3), the FAS (death receptor)-mediated initiator caspases (including caspase 8), and stress-induced caspases (including caspase 9), were tested for their protective effects. Inhibition of caspases 3, 8, and 9 each robustly protected OLs following insult with staurosporine, thapsigargin, or kainate when added at optimal times. The time of addition of the inhibitors for maximal protection varied with the agent, from 1 h of preincubation before addition of staurosporine to 6 h after addition of kainate. Much less protection was seen for the NO generator SNAP under any condition. The role of calcium in OL death in each model was investigated by chelating extracellular Ca++ with EGTA, and by inhibiting the Ca++-activated calpain proteases. Calcium chelation did not protect against staurosporine, but decreased OL death initiated by kainate, thapsigargin, or NO. The calpain inhibitors PD150606 and calpain inhibitor I protected from cell death initiated by staurosporine, kainate, and thapsigargin, but not from cell death initiated by the NO donor SNAP.  相似文献   

17.
18.
Electrical stimulation of the cerebellar fastigial nucleus (FN) elicits a prolonged ( approximately 10 days) and substantial (50-80%) protection against ischemic and excitotoxic injuries. The mechanism(s) of protection are unknown. We investigated whether FN stimulation directly protects brain cells against apoptotic cell death in an in vitro rat brain slice culture model. Rats were electrically stimulated in FN or, as control, the cerebellar dentate nucleus (DN). Coronal slices through the forebrain were explanted, exposed to staurosporine, harvested, and analyzed for caspase-3 activity by a fluorescence assay. FN, but not DN, stimulation significantly reduced staurosporine-induced caspase-3 activity by 39 +/- 7% at 3 h, 31 +/- 3% at 6 h and 26 +/- 4% at 10 h of incubation. Immunocytochemistry revealed FN-specific reductions in activated caspase-3 mainly in glial-like cells throughout the forebrain. FN stimulation also results in a 56.5% reduction in cytochrome c release upon staurosporine incubation. We conclude that neuroprotection elicited from FN stimulation can directly modify the sensitivity of brain cells to apoptotic stimuli and thereby suppress staurosporine induced apoptosis in adult rat brain slices. This model indicates that neuroprotection can be studied in vitro and provides new insight into the potential role of glial cells in ischemic protection of neurons induced by FN stimulation.  相似文献   

19.
The T cell Ag (Ti-CD3) receptor complex has been proposed to regulate phosphoinositide-specific phospholipase C (PLC) through a cholera toxin (CTX)-sensitive guanine nucleotide-binding (G) protein. In this study, we have used CTX and staurosporine as pharmacologic probes to further define the linkage between the Ti-CD3 receptor and PLC activity in the human T cell line, Jurkat. CTX pretreatment inhibited Ti-CD3 receptor-dependent phosphoinositide hydrolysis and, concomitantly, protein tyrosine kinase activation in intact cells. Studies with electrically permeabilized Jurkat cells revealed that guanosine 5'-(3-O-thio) triphosphate stimulated an increase in PLC activity, that unlike the response to Ti-CD3 receptor ligation, was not affected by cellular pretreatment with CTX. In contrast, the phosphotyrosine phosphatase inhibitors, orthovanadate and molybdate anions, stimulated phosphoinositide hydrolysis in permeabilized cells through a CTX-sensitive mechanism of PLC activation. Additional studies with a known PTK inhibitor, staurosporine, supported the results obtained with CTX. Staurosporine pretreatment inhibited the phosphoinositide hydrolysis induced by anti-CD3 antibodies or phosphotyrosine phosphatase inhibitors, but failed to alter the G protein-dependent PLC activation response to guanosine 5'-(3-O-thio) triphosphate. The results of this study indicate that PLC activity(s) in Jurkat cells are regulated by both G protein- and PTK-dependent coupling mechanisms. However, the differential inhibitory effects of CTX and staurosporine on these PLC activation pathways strongly suggest that a protein tyrosine kinase activation event, rather than a G protein, mediates the functional linkage between the Ti-CD3 receptor and PLC activity in Jurkat cells.  相似文献   

20.
Full activation of protein kinase B (PKB, also called Akt) requires phosphorylation on two regulatory sites, Thr-308 in the activation loop and Ser-473 in the hydrophobic C-terminal regulatory domain (numbering for PKB alpha/Akt-1). Although 3'-phosphoinositide-dependent protein kinase 1 (PDK1) has now been identified as the Thr-308 kinase, the mechanism of the Ser-473 phosphorylation remains controversial. As a step to further characterize the Ser-473 kinase, we examined the effects of a range of protein kinase inhibitors on the activation and phosphorylation of PKB. We found that staurosporine, a broad-specificity kinase inhibitor and inducer of cell apoptosis, attenuated PKB activation exclusively through the inhibition of Thr-308 phosphorylation, with Ser-473 phosphorylation unaffected. The increase in Thr-308 phosphorylation because of overexpression of PDK1 was also inhibited by staurosporine. We further show that staurosporine (CGP 39360) potently inhibited PDK1 activity in vitro with an IC(50) of approximately 0.22 microm. These data indicate that agonist-induced phosphorylation of Ser-473 of PKB is independent of PDK1 or PKB activity and occurs through a distinct Ser-473 kinase that is not inhibited by staurosporine. Moreover, our results suggest that inhibition of PKB signaling is involved in the proapoptotic action of staurosporine.  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