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 15 毫秒
1.
A protein kinase C (PKC) activating factor (AF) has been identified in the extracellular medium of V3.17 vincristine resistant murine erythroleukemia (MEL) cells clone. The factor is a protein that stimulates the activity of PKC alpha and beta isozymes isolated from MEL cells, rat and mouse brain approximately 2 to 2.5 fold over the Vmax, respectively. AF promotes an identical activation in the presence of all the effectors but also when the amount of Ca2+ is reduced to microM concentration and in the absence of diacylglycerol (DAG). The factor shows a greater activating efficiency with PKC beta isozymes. AF binds to PKC presumably at the DAG binding site as suggested by the competition between phorbol dibutyrate and AF for binding to the kinase. Moreover, AF promotes the selective binding of PKC beta to natural or artificial membranes in the presence of microM concentrations of Ca2+. Altogether these results suggest the presence in MEL cells of a protein factor that can promote association of PKC to the membranes together with activation of the kinase, without the requirement for DAG formation. This could be visualized as a new mechanism for prolonged and selective activation of PKC.  相似文献   

2.
Mechanisms of Ca2+ sensitization of both myosin light chain (MLC) phosphorylation and force development by protein kinase C (PKC) were studied in permeabilized tonic smooth muscle obtained from the rabbit femoral artery. For comparison, the Ca2+ sensitizing effect of guanosine 5'-O-(gamma-thiotriphosphate) (GTP gamma S) was examined, which had been previously shown to inhibit MLC phosphatase in phasic vascular smooth muscle. We now report that PKC activators (phorbol esters, short chain synthetic diacylglycerols and a diacylglycerol kinase inhibitor) and GTP gamma S significantly increase both MLC phosphorylation and force development at constant [Ca2+]. Major phosphorylation site occurring in the presence of phorbol-12,13- dibutyrate (PDBu) or GTP gamma S at constant [Ca2+] is the same serine residue (Ser-19) as that phosphorylated by MLC kinase in response to increased Ca2+ concentrations. In an ATP- and Ca(2+)-free solution containing 1-(5-chloronaphthalene-1-sulfonyl)-1H-hexahydro-1,4- diazepine (ML-9), to avoid the kinase activity, both PDBu and GTP gamma S significantly decreased the rate of MLC dephosphorylation to half its control value. However, PDBu inhibited the relaxation rate more than did GTP gamma S. In the presence of microcystin-LR to inhibit the phosphatase activity, neither PDBu nor GTP gamma S affected MLC phosphorylation and force development. These results indicate that PKC, like activation of GTP binding protein, increases Ca2+ sensitivity of both MLC phosphorylation and force production through inhibition of MLC phosphatase.  相似文献   

3.
C A O'Brian  N E Ward 《Biochemistry》1991,30(9):2549-2554
We recently reported that autophosphorylated rat brain protein kinase C (PKC) catalyzes a Ca2(+)- and phosphatidylserine- (PS-) dependent ATPase reaction. The Ca2(+)- and PS-dependent ATPase and histone kinase reactions of PKC each had a Km app(ATP) of 6 microM. Remarkably, the catalytic fragment of PKC lacked detectable ATPase activity. In this paper, we show that subsaturating concentrations of protein substrates accelerate the ATPase reaction catalyzed by PKC and that protein and peptide substrates of PKC induce ATPase catalysis by the catalytic fragment. At subsaturating concentrations, histone III-S and protamine sulfate each accelerated the ATPase activity of PKC in the presence of Ca2+ and PS by as much as 1.5-fold. At saturating concentrations, the protein substrates were inhibitory. Poly(L-lysine) failed to accelerate the ATPase activity, indicating that the acceleration observed with histone III-S and protamine sulfate was not simply a result of their gross physical properties. Furthermore, histone III-S induced the ATPase activity of the catalytic fragment of PKC, at both subsaturating and saturating histone concentrations. The induction of ATPase activity was also elicited by the peptide substrate Arg-Arg-Lys-Ala-Ser-Gly-Pro-Pro-Val, when the peptide was present at concentrations near its Km app. The induction of the ATPase activity by the nonapeptide provides strong evidence that the binding of phospho acceptor substrates to the active site of PKC can stimulate ATP hydrolysis. Taken together, our results indicate that PKC-catalyzed protein phosphorylation is inefficient, since it is accompanied by Pi production.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
The ability of Ca2(+)-mobilizing hormones to promote changes in the subcellular distribution of protein kinase C (PKC) was studied in isolated hepatocytes. In recently isolated cells the distribution of PKC between the soluble and particulate fractions was 47 and 53% respectively. Exposure of the hepatocytes to 100 nM-vasopressin produced an increased phosphoinositide turnover, as reflected by the changes in the concentrations of inositol trisphosphate and Ca2+, and in glycogen phosphorylase a activity. However, the distribution of both PKC activity and [3H]phorbol dibutyrate binding between the cytosol and the membranes remained unchanged under these conditions. To determine the threshold values of the concentrations of Ca2+ and diacylglycerol required to produce a redistribution of PKC, the hepatocytes were treated with the Ca2+ ionophore ionomycin, and with permeant diacylglycerol derivatives. Hepatocytes incubated in the presence of 100 nM-vasopressin required concentrations of Ca2+ 2.5 times those produced physiologically by the hormone to produce translocation of PKC from the cytosol to the membranes. These studies suggest that, at least in hepatocytes, activation of PKC in response to Ca2(+)-mobilizing hormones involves only the pre-existent membrane-bound enzyme without affecting the soluble enzyme.  相似文献   

5.
The effects of a synthetic protein kinase C (PKC) activator, (-)-indolactam V (ILV), were studied in SH-SY5Y human neuroblastoma cells. (-)-ILV induced a translocation of PKC from cytosol to plasma membrane and displaced 3H-phorbol dibutyrate binding in the micromolar range. In addition, (-)-ILV caused a decreased sensitivity of cells to muscarinic agonist-induced Ca2+ mobilization measured with quin-2 and induced a down-regulation of cell surface muscarinic receptors. All the changes induced by (-)-ILV were similar in magnitude to those seen with the phorbol ester tetradecanoyl phorbol acetate (TPA). The results suggest that (-)-ILV is a full activator of PKC and a promising alternative to phorbol esters in studies on mechanism of actions of PKC.  相似文献   

6.
A mixed micellar assay for the binding of phorbol-esters to protein kinase C was developed to investigate the specificity and stoichiometry of phospholipid cofactor dependence and oligomeric state of protein kinase C (Ca2+/phospholipid-dependent enzyme) required for phorbol ester binding. [3H]Phorbol dibutyrate was bound to protein kinase C in the presence of Triton X-100 mixed micelles containing 20 mol % phosphatidylserine (PS) in a calcium-dependent manner with a Kd of 5 X 10(-9) M. The [3H]phorbol dibutyrate X protein kinase C . Triton X-100 . PS mixed micellar complex eluted on a Sephacryl S-200 molecular sieve at an Mr of approximately 200,000; this demonstrates that monomeric protein kinase C binds phorbol dibutyrate. This conclusion was supported by molecular sieve chromatography of a similar complex where Triton X-100 was replaced with beta-octylglucoside. Phorbol dibutyrate activation of protein kinase C in Triton X-100/PS mixed micelles occurred and was dependent on calcium. The PS dependence of both phorbol ester activation and binding to protein kinase C lagged initially and then was highly cooperative. The minimal mole per cent PS required was strongly dependent on the concentration of phorbol dibutyrate or phorbol myristic acetate employed. Even at the highest concentration of phorbol ester tested, a minimum of 3 mol % PS was required; this indicates that approximately four molecules of PS are required. [3H]Phorbol dibutyrate binding was independent of micelle number at 20 mol % PS. The phospholipid dependencies of phorbol ester binding and activation were similar, with PS being the most effective; anionic phospholipids (cardiolipin, phosphatidic acid, and phosphatidylglycerol were less effective, whereas phosphatidylcholine, phosphatidylethanolamine, and sphingomyelin did not support binding or activation. sn-1,2-Dioleoylglycerol displaced [3H]phorbol dibutyrate quantitatively and competitively. The data are discussed in relation to a molecular model of protein kinase C activation.  相似文献   

7.
The effects of agents that modulate intracellular release of calcium and protein kinase C (PKC) activation on noradrenaline (NA)-induced contractions of epididymal vas deferens in calcium-free/EGTA (1 mM) medium were investigated. NA (100 microM) or methoxamine (100 microM) evoked repeatable contractions. Clonidine (100-300 microM) was ineffective. The contractions to NA were reduced by procaine (1-10 mM) but not by thapsigargin (0.1-30 microM), ryanodine (1-30 microM) or TMB-8 (1-30 microM). Contractions to cumulative additions of NA (1-100 microM) were enhanced in the presence of cyclopiazonic acid (10 & 30 microM) but not ryanodine (10 & 30 microM). Sequential contractions to NA were not blocked by PKC inhibitors, calphostin C (1 microM) or Ro 31-8220 (1-30 microM) but were reduced by H-7 (1-30 microM), a broad spectrum protein kinase inhibitor. Although RT-PCR experiments detected mRNA for some Ca2+-dependent/DAG-activated and Ca2+-independent/DAG-activated PKC isoforms in epididymal vas deferens, the PKC activators, phorbol 12, 13-dibutyrate (100 microM) or phorbol 12-myristate 13-acetate (100 microM) failed to activate the tissues in calcium-free medium but enhanced subsequent contractions to NA. These results indicate a limited role for intracellular calcium stores and phorbol ester/DAG-sensitive PKC isoforms in NA-induced contraction of epididymal rat vas deferens in calcium-free medium. The results suggest that pharmacomechanical coupling triggered by NA may involve the sensitization of contractile myofilaments to Ca2+ or a Ca2+-independent mechanism. The possible involvement of Ca2+-independent/DAG-insensitive PKC isoforms and agonist-dependent but PKC-independent sensitization pathway is discussed.  相似文献   

8.
The phenolic antioxidant 2,6-bis(1,1-dimethyl ethyl)-4-methylphenol (BHT) evokes a transient phosphorylation of two platelet proteins of Mr 20,000 and 47,000 that are well-known substrates of protein kinase C (PKC) and, similarly to phorbol esters, a slight but persistent phosphorylation of a protein of Mr 26,000. These effects are observed both in the presence and in the absence of extracellular calcium, but are abolished in the presence of the protein kinase C inhibitor staurosporine. The phosphorylation of the 47 kDa protein takes place mostly at the serine and, to a lesser extent, at threonine residues. BHT induces an increased binding of tritiated phorbol dibutyrate to platelets indicating a PKC translocation from cytosol to plasma membrane. Addition of BHT (20 microM) a few min prior to thrombin causes inhibition of both agonist-evoked protein phosphorylation and increase in the Ca2+ concentration, the latter inhibition being counteracted by staurosporine. The inhibitory effect lasts for several minutes even after removal of BHT from the cellular suspending medium. Similar results are obtained with nordihydroguaiaretic acid, whereas 2- and 3-tert-butyl-4-methoxyphenol (BHA) produce only slight effects. BHT activates the protein kinase C purified from pig brain in a concentration-dependent manner (up to 200 microM), whereas it does not affect the activity of other purified protein kinases such as type 1 and 2 casein kinases, type II A, II B and III tyrosine protein kinases from rat spleen and the catalytic subunit of cyclic AMP-dependent protein kinase. It is concluded that, similarly to diacylglycerols and phorbol esters, these phenolic antioxidants activate the protein kinase C, which in turn desensitizes platelets towards subsequent phospholipase C activation.  相似文献   

9.
Both protein kinase C and Ca2+ may act in concert to bring about activation of secretion. This study examined the actions on pancreatic acini of ionomycin and phorbol dibutyrate, which selectively stimulate one or the other of these pathways; their stimulatory effects were compared with those of receptor agonists, such as carbachol and caerulein, which activate phospholipase C. The Ca2+ ionophore ionomycin produced a dose-dependent increase in amylase secretion and intracellular free Ca2+ (as measured by quin-2). The increase in amylase secretion elicited by carbachol or caerulein was accompanied by a small sustained increase in intracellular free Ca2+, following an initial peak. However, the elevation in intracellular free Ca2+ produced by these receptor agonists for a given level of amylase secretion was less than that observed with ionomycin. Phorbol dibutyrate stimulated amylase secretion by a mechanism that was independent of extracellular Ca2+, and no change in intracellular free Ca2+ was observed. Synergistic stimulatory effects of phorbol dibutyrate and ionomycin were observed, whether the phorbol ester was present before, or in combination with, ionomycin. Diacylglycerols containing unsaturated fatty acids (1,2-dioleoylglycerol and 1,3-dioleoylglycerol) also stimulated amylase secretion and exhibited synergistic effects on secretion with ionomycin. These findings suggest that complete activation of amylase secretion from the pancreas requires stimulation of both Ca2+-dependent and protein kinase C-activated pathways.  相似文献   

10.
Interaction of protein kinase C (PKC) isozymes with phosphatidylinositol 4,5-bisphosphate (PIP2) was investigated by monitoring the changes in the intrinsic fluorescence of the enzyme, the kinase activity, and phorbol ester binding. Incubation of PKC I, II, and III with PIP2 resulted in different rates of quenching of PKC fluorescence and different degrees of inactivation of these enzymes. Other inositol-containing phospholipids such as phosphatidylinositol and phosphatidylinositol 4-phosphate also caused differential rates of quenching of the intrinsic fluorescence of these enzymes. These latter two phospholipids were, however, less potent in the inactivation of PKCs than PIP2. The IC50 of PIP2 were 2, 4, and 11 microM for PKC I, II, and III, respectively. Inactivation of PKCs by PIP2 cannot be reversed by extensive dilution of PIP2 with Nonidet P-40 nor by digestion of PIP2 with phospholipase C. Interaction of PIP2 with the various PKC isozymes was greatly facilitated in the presence of Mg2+ or Ca2+ as evidenced by the accelerated quenching of the PKC fluorescence, however, these divalent metal ions protected PKC from the PIP2-induced inactivation. Binding of PIP2 to PKC in the absence of divalent metal ion also caused a reduction of [3H]phorbol 12,13-dibutyrate binding as a result of reducing the affinity of the enzyme for phorbol ester. Based on gel filtration chromatography, it was estimated that one molecule of PKC interacted with one PIP2 micelle with an aggregation number of 80-90. The PIP2-bound PKC could further interact with phosphatidylserine in the presence of Ca2+ to form a larger complex. Binding of PKC to both PIP2 and phosphatidylserine in the presence of Ca2+ was also evident by changes in the intrinsic fluorescence of PKC. As the interaction of PKC with PIP2, but not with phosphatidylserine, could be enhanced by millimolar concentrations of Mg2+, we propose that PIP2 may be a component of the membrane anchor for PKC under basal physiological conditions when [Ca2+]i is low and Mg2+ is plentiful. Under the in vitro assay conditions, PIP2 could stimulate PKC activity to a level approximately 10-20% of that by diacylglycerol. The stimulatory effect of PIP2 on PKC apparently is not due to binding to the same site recognized by diacylglycerol or phorbol ester, because PIP2 cannot effectively compete with phorbol 12,13-dibutyrate in the binding assay.  相似文献   

11.
We have constructed the expression plasmids harboring protein kinase C (PKC) mutant cDNAs with a series of deletions in the PKC coding region. These plasmids were transfected into COS7 cells to characterize the PKC mutants. Immunoblot analysis using the anti-PKC antibody identified proteins with the Mr values expected from the PKC mutant cDNAs in the extracts from COS7 cells. The wild-type PKC, when expressed in COS7 cells, conferred increased phorbol ester binding activity on intact cells; but the PKC mutants with the deletion around the C1 region did not show this activity. The wild-type PKC showed protein kinase activity dependent on phospholipid, Ca2+, and phorbol ester, whereas these PKC mutants exhibited protein kinase activity independent of the activators in a cell-free system. A PKC mutant cDNA with the deletion in the C2 region gave increased phorbol ester binding activity. Protein kinase activity of this mutant was much less dependent on Ca2+ compared with the wild-type PKC. A PKC mutant cDNA with the deletion in the C3 region conferred increased phorbol ester binding activity, but neither activator-dependent nor -independent protein kinase activity. These results indicate that elimination of the C1 region of PKC gives rise to constitutively active PKC independent of phospholipid, Ca2+, and phorbol ester and that the C1-C3 regions play distinct roles in the regulatory and catalytic function of PKC. In another series of experiments, transfection of some PKC mutant cDNAs with the deletions around the C1 region into Chinese hamster ovary and Jurkat cells activated the activator protein-1-binding element or the c-fos gene enhancer linked to the chloramphenicol acetyltransferase reporter gene in the absence of phorbol ester. Microinjection of these constructs into Xenopus oocytes induced initiation of germinal vesicle breakdown, indicating that they stimulated the PKC pathway in vivo. Thus, the phorbol ester-independent PKC mutant cDNAs could be a powerful tool to investigate the transmembrane signaling pathway mediated by PKC.  相似文献   

12.
There are specified and saturable binding sites for [20-3H]phorbol-12,13-dibutyrate on enzymatically dissociated rat cardiac myocytes. At 37 degrees C, maximal binding occurs within 20 min, with a KD of 3.9 nM and Bmax of 0.275 pmol/mg. [3H]Phorbol dibutyrate binding is blocked by 12-O-tetradecanoyl phorbol-13-acetate but not by 4 alpha-phorbol or 4 alpha-phorbol-12,13-dibutyrate. Dibucaine, tetracaine, chlorpromazine, and phospholipase C lowered phorbol binding through a competitive mechanism. Similarly, unsaturated (but not saturated) diacylglycerols competed with [3H]phorbol dibutyrate for the binding site. There was a progressive decline in specific binding of phorbol diesters to cardiac myocytes which occurred primarily during the first 3 weeks of postnatal life. Cardiac phorbol diester receptors may mediate protein kinase C-dependent effects on important cellular functions such as Ca2+ transport.  相似文献   

13.
Clonal GH4C1 rat pituitary cells are heterogeneous with respect to phorbol dibutyrate receptors (PDBu-R) and protein kinase C (PKC) content. GH cell PDBu-Rs can be separated into two categories based on Ca2(+)-modulation of receptor affinity. Approximately 70% of the cytosolic PDBu-Rs demonstrate Ca2(+)-sensitive receptor affinity and redistribute from the soluble to the particulate fraction in the presence of excess Ca2+. The other 30% of the receptors remain in the cytosol in the presence of excess Ca2+. Their receptor affinity is Ca2(+)-independent. Northern blot hybridization and immunoblot analysis showed that GH4C1 cells express Ca2(+)-independent epsilon-PKC as well as Ca2(+)-dependent alpha- and beta-PKCs. Cell lysis in Ca2+ caused the redistribution of greater than 95% of alpha- and beta-PKC to the particulate fraction, whereas approximately 90% of the epsilon-PKC remained in the cytosol. In contrast, brief treatment of GH cell cultures with PDBu or thyrotropin-releasing hormone caused redistribution of all three isozymes. Prolonged treatment with PDBu down-modulated all three isozymes but at different rates and to different extents. In contrast, prolonged thyrotropin-releasing hormone treatment selectively down-modulated epsilon-PKC. These results demonstrate that GH cells have both Ca2(+)-sensitive and -insensitive PKCs and PDBu-Rs and that both populations are regulated by agonists that control prolactin synthesis and secretion by these cells.  相似文献   

14.
Sphingosine inhibited protein kinase C activity and phorbol dibutyrate binding. When the mechanism of inhibition of activity and phorbol dibutyrate binding was investigated in vitro using Triton X-100 mixed micellar methods, sphingosine inhibition was subject to surface dilution; 50% inhibition occurred when sphingosine was equimolar with sn-1,2-dioleoylglycerol (diC18:1) or 40% of the phosphatidylserine (PS) present. Sphingosine inhibition was modulated by Ca2+ and by the mole percent of diC18:1 and PS present. Sphingosine was a competitive inhibitor with respect to diC18:1, phorbol dibutyrate, and Ca2+. Increasing levels of PS markedly reduced inhibition by sphingosine. Since protein kinase C activity shows a cooperative dependence on PS, the kinetic analysis of competitive inhibition was only suggestive. Sphingosine inhibited phorbol dibutyrate binding to protein kinase C but did not cause protein kinase C to dissociate from the mixed micelle surface. Sphingosine addition to human platelets blocked thrombin and sn-1,2-dioctanoylglycerol-dependent phosphorylation of the 40-kDa (47 kDa) dalton protein. Moreover, sphingosine was subject to surface dilution in platelets. The mechanism of sphingosine inhibition is discussed in relation to a previously proposed model of protein kinase C activation. The possible physiological role of sphingosine as a negative effector of protein kinase C is suggested and a plausible cycle for its generation is presented. The potential physiological significance of sphingosine inhibition of protein kinase C is further established in accompanying papers on HL-60 cells (Merrill, A. H., Jr., Sereni, A. M., Stevens, V. L., Hannun, Y. A., Bell, R. M., Kinkade, J. M., Jr. (1986) J. Biol. Chem. 261, 12010-12615) and human neutrophils (Wilson, E., Olcott, M. C., Bell, R. M., Merrill, A. H., Jr., and Lambeth, J. D. (1986) J. Biol. Chem. 261, 12616-12623). These results also suggest that sphingosine will be a useful inhibitor for investigating the function of protein kinase C in vitro and in living cells.  相似文献   

15.
Calcium phospholipid dependent protein kinase C (PKC) is activated by diacylglycerol (DG) and by phorbol esters and is recognized to be the phorbol ester receptor of cells; DG displaces phorbol ester competitively from PKC. A phospholipid, phosphatidylinositol 4,5-bisphosphate (PIP2), can also activate PKC in the presence of phosphatidylserine (PS) and Ca2+ with a KPIP2 of 0.04 mol %. Preliminary experiments have suggested a common binding site for PIP2 and DG on PKC. Here, we investigate the effect of PIP2 on phorbol ester binding to PKC in a mixed micellar assay. In the presence of 20 mol % PS, PIP2 inhibited specific binding of [3H]phorbol 12,13-dibutyrate (PDBu) in a dose-dependent fashion up to 85% at 1 mol %. Inhibition of binding was more pronounced with PIP2 than with DG. Scatchard analysis indicated that the decrease in binding of PDBu in the presence of PIP2 is the result of an altered affinity for the phorbol ester rather than of a change in maximal binding. The plot of apparent dissociation constants (Kd') against PIP2 concentration was linear over a range of 0.01-1 mol % with a Ki of 0.043 mol % and confirmed the competitive nature of inhibition between PDBu and PIP2. Competition between PIP2 and phorbol ester could be demonstrated in a liposomal assay system also. These results indicate that PIP2, DG, and phorbol ester all compete for the same activator-receiving region on the regulatory moiety of protein kinase C, and they lend support to the suggestion that PIP2 is a primary activator of the enzyme.  相似文献   

16.
Phorbol ester-induced conventional protein kinase C (PKCalpha, -betaIota/IotaIota, and -gamma) isozyme activities are potentiated by 1,2-diacyl-sn-glycerol. This has been attributed to a "cooperative" interaction of the two activators with two discrete sites termed the low- and high-affinity phorbol ester binding sites, respectively [Slater, S. J., Milano, S. K., Stagliano, B. A., Gergich, K. J., Ho, C., Mazurek, A., Taddeo, F. J., Kelly, M. B., Yeager, M. D., and Stubbs, C. D. (1999) Biochemistry 38, 3804-3815]. Here, we report that the 1-O-alkyl ether diglyceride, 1-O-hexadecyl-2-acetyl-sn-glycerol (HAG), like its 1,2-diacyl counterpart, 1-oleoyl-2-acetyl-sn-glycerol (OAG), also potentiated PKCalpha, -betaI/II, and -gamma activities induced by the phorbol ester 4beta-12-O-tetradecanoylphorbol-13-acetate (TPA). Similar to OAG, HAG was found to bind to the low-affinity phorbol ester binding site and to enhance high-affinity phorbol ester binding, and to decrease the level of Ca(2+) required for phorbol ester-induced activity, while being without effect on the Ca(2+) dependence of membrane association. Thus, similar to OAG, HAG may also potentiate phorbol ester-induced activity by interacting with the low-affinity phorbol ester binding site, leading to a reduced level of Ca(2+) required for the activating conformational change. However, HAG was found not to behave like a 1,2-diacyl-sn-glycerol in that alone it did not induce PKC activity, and also in that it enhanced OAG-induced activity. The results reveal HAG to be a member of a new class of "nonactivating" compounds that modulate PKC activity by interacting with the low-affinity phorbol ester binding site.  相似文献   

17.
The receptor for tumor-promoting phorbol esters has been shown to be the Ca+2/phospholipid dependent enzyme protein kinase C (PKC). There are two major groups of PKC, the conventional PKC isotypes alpha, beta I, beta II, gamma) and the novel Ca+2-independent PKC (delta, epsilon, zeta, eta). Phorbol esters previously have been demonstrated to increase human IFN-gamma gene expression after treatment of a murine T cell line (Cl 9) that has been transfected with human IFN-gamma genomic DNA. In contrast, treatment with Ca+2 ionophore alone or in combination with phorbol ester did not enhance IFN-gamma production in a synergistic manner above the level obtained with phorbol ester treatment alone. To determine whether the lack of effect of Ca+2 ionophore is due to a defect in PKC, we compared the level of PKC autophosphorylation in the mouse T cell line (Cl 9), a mouse epidermal cell line (JB6), and purified rat brain PKC by in vitro kinase assays. The results demonstrate that instead of the expected 80-kDa autophosphorylated PKC band seen in purified rat brain PKC or mouse JB6 cell lysates, only a novel 97-kDa Ca+2-independent phosphoprotein was observed in Cl 9 cells. To ascertain if there was any nucleic acid sequence similarity to PKC epsilon, we hybridized Cl 9 poly(A+) RNA with a cloned fragment of the PKC epsilon gene and observed two hybridizing RNA bands (4.4 and 4.0 kb). Our results suggest that the 97-kDa phosphoprotein is similar to, but not identical with, PKC epsilon and is the major PKC expressed in the Cl 9 murine T cell line. These data suggested that the 97-kDa PKC may be responsible for the induction of both the transfected human IFN-gamma gene and the endogenous murine IL-2R alpha-chain.  相似文献   

18.
Suramin inhibited protein kinase C (PKC) type I-III activity in a concentration-dependent manner. Similar inhibitory effects were observed with M-kinase, the constitutively active catalytic fragment of PKC, and autophosphorylation of PKC types I-III. Kinetic experiments indicated that suramin competitively inhibits activity with respect to ATP (Ki = 17, 27, and 31 microM, respectively) and that it can also inhibit by interaction with the substrate histone III-S. With protamine as the Pi acceptor, suramin inhibition was dependent on lipid, being approximately 4-fold less sensitive to inhibition in the absence of phosphatidylserine and diacylglycerol than in their presence. Suramin at low concentrations (10-40 microM), in the presence of Ca2+ and absence of lipid, was able to stimulate kinase activity (approximately 200-400%) in a type-dependent manner and at higher concentrations inhibited activity with histone III-S as substrate. These results indicate that suramin, a hexa-anionic hydrophobic compound, can act as a negatively charged phospholipid analog in activating PKC in the presence of Ca2+ and absence of lipid and can inhibit Ca2+/phosphatidylserine/diacylglycerol-stimulated kinase activity at higher concentrations by competing with ATP or by interaction with the exogenous substrate. Suramin inhibited cAMP-dependent protein kinase much less potently (IC50 = 656 microM) than PKC. The ability of suramin to inhibit PKC-mediated processes in intact cells was tested using the phorbol ester-stimulated respiratory burst of neutrophils as a model system. The respiratory burst of human neutrophils, when preincubated with suramin and then stimulated with phorbol ester, was inhibited in a concentration-dependent manner, suggesting that suramin may also be able to inhibit PKC-mediated processes in intact cells.  相似文献   

19.
Protein kinase C (PKC) molecular species of GH4C1 cells were analyzed after separation by hydroxyapatite column chromatography. A novel Ca2(+)-independent PKC, nPKC epsilon, was identified together with two conventional Ca2(+)-dependent PKCs, PKC alpha and beta II by analysis of kinase and phorbol ester-binding activities, immunoblotting using isozyme-specific antibodies, and Northern blotting. These PKCs are down-regulated differently when cells are stimulated by outer stimuli; phorbol esters deplete PKC beta II and nPKC epsilon from the cells more rapidly than PKC alpha, whereas thyrotropin-releasing hormone (TRH) at 200 nM depletes nPKC epsilon exclusively with a time course similar to that induced by phorbol esters. However, translocation of PKC alpha and beta II to the membranes is elicited by both TRH and phorbol esters. These results suggest that TRH and phorbol ester activate PKC alpha and beta II differently but that nPKC epsilon is stimulated similarly by both stimuli. Thus, in GH4C1 cells, Ca2(+)-independent nPKC epsilon may play a crucial role distinct from that mediated by Ca2(+)-dependent PKC alpha and beta II in a cellular response elicited by both TRH and phorbol esters.  相似文献   

20.
Electrically permeabilized RINm5F cells were used to assess the factors required for activation of protein kinase C (PKC) and insulin secretion. PKC was activated either by phorbol 12-myristate 13-acetate (PMA) or by the generation of endogenous diacylglycerol in response to the nonhydrolyzable guanine nucleotide analog guanosine 5'-O-(thiotriphosphate) (GTP gamma S). As shown previously, both PMA and GTP gamma S elicit Ca2+-independent insulin secretion. This effect was mimicked by guanyl-5'-yl imidodiphosphate (Gpp(NH)p) but not by guanosine 5'-O-(3-fluorotriphosphate) and guanosine 5'-O-(3-phenyltriphosphate) possessing only one negative charge in the gamma-phosphate group. The action of PMA was mediated by PKC, since the agent caused both phosphorylation of specific protein substrates and association of the enzyme with cellular membranes. This translocation was independent of the Ca2+ concentration employed. In contrast, GTP gamma S only promoted association of PKC with membranes at 10(-6) and 10(-5) M Ca2+ and failed to alter significantly protein phosphorylation in the absence of Ca2+. Neither Gpp(NH)p, which stimulates insulin release, nor the other two GTP analogs, increased the proportion of PKC associated with membranes. To verify that the Ca2+-dependent effect of GTP gamma S on PKC is due to activation of phospholipase C, we measured the generation of diacylglycerol. GTP gamma S indeed stimulated diacylglycerol production in the leaky cells by about 50% at Ca2+ concentrations between 10(-7) and 10(-5) M, an effect which was almost abolished in the absence of Ca2+. Thus, at 10(-7) M Ca2+, the concentration found in resting intact cells, the generated diacylglycerol was not sufficient to cause PKC insertion into the membrane, demonstrating that both elevated Ca2+ and diacylglycerol are necessary for translocation to occur. It is concluded that while PKC activation by PMA elicits Ca2+-independent insulin secretion, the kinase seems not to mediate the stimulatory action of GTP analogs in the absence of Ca2+.  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