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 125 毫秒
1.
Immunochemical characterization of rat brain protein kinase C   总被引:11,自引:0,他引:11  
Polyclonal antibodies against rat brain protein kinase C (the Ca2+/phospholipid-dependent enzyme) were raised in goat. These antibodies can neutralize completely the kinase activity in purified enzyme preparation as well as that in the crude homogenate. Immunoblot analysis of the purified and the crude protein kinase C preparations revealed a major immunoreactive band of 80 kDa. The antibodies also recognize the same enzyme from other rat tissues. Neuronal tissues (cerebral cortex, cerebellum, hypothalamus, and retina) and lymphoid organs (thymus and spleen) were found to be enriched in protein kinase C, whereas lung, kidney, liver, heart, and skeletal muscle contained relatively low amounts of this kinase. Limited proteolysis of the purified rat brain protein kinase C with trypsin results in an initial degradation of the kinase into two major fragments of 48 and 38 kDa. Both fragments are recognized by the antibodies. However, further digestion of the 48-kDa fragment to 45 kDa and the 38-kDa fragment to 33 kDa causes a loss of the immunoreactivity. Upon incubation of the cerebellar extract with Ca2+, the 48-kDa fragment was also identified as a major proteolytic product of protein kinase C. Proteolytic degradation of protein kinase C converts the Ca2+/phospholipid-dependent kinase to an independent form without causing a large impairment of the binding of [3H]phorbol 12,13-dibutyrate. The two major proteolytic fragments were separated by ion exchange chromatography and one of them (45-48 kDa) was identified as a protein kinase and the other (33-38 kDa) as a phorbol ester-binding protein. This degraded form of the phorbol ester-binding protein still requires phospholipid for activity but, unlike the native enzyme, becomes less dependent on Ca2+. These results demonstrate that rat brain protein kinase C is composed of two functionally distinct units, namely, a protein kinase and a Ca2+-independent/phospholipid-dependent phorbol ester-binding protein.  相似文献   

2.
We have examined the immunocytochemical localization of protein kinase C (PKC) in NIH 3T3 cells using mAbs that recognize Type 3 PKC. In control cells, the immunofluorescent staining was similar with mAbs directed to either the catalytic or the regulatory domain of PKC. Type 3 PKC localized in a diffuse cytoplasmic pattern, while the nuclei were apparently unstained. Cytoskeletal components also were Treatment of the cells with phorbol 12-myristate 13-acetate (PMA) resulted in a redistribution of PKC with a specific increase in nuclear PKC. Compared to control cells, the staining with the anticatalytic domain mAbs changed markedly, covering the entire cell surface. In contrast, the staining by the antiregulatory domain mAb did not cover the cell surface and the nuclei remained unstained; these results suggest that PKC activation leads to a conformational change of the regulatory domain such that the epitope recognized by the antiregulatory domain mAb is not readily accessible. We have demonstrated by three criteria that PMA treatment specifically increased PKC in the nucleus: (a) immunofluorescent staining in isolated nuclei increased; (b) Western blots showed that our mAbs detected only one protein, the 82-kD PKC, whose level increased in nuclear lysates from PMA-treated cells; and (c) PKC activity increased in nuclear lysates. In fractionation studies we demonstrated that PKC specifically localized to the nuclear envelope fraction. These results demonstrate that PMA activation leads to a rapid redistribution of Type 3 PKC to the nuclear envelope, and suggests that this isozyme may play a role in mediating PKC-induced changes in gene expression.  相似文献   

3.
Trypsinization of rat brain protein kinase C (80 kDa) into 50- and 32-kDa fragments occurred without inhibition of [3H]phorbol dibutyrate ([3H]PDBu) binding activity. The 50-kDa fragment, the catalytic domain (Inoue, M., Kishimoto, A., Takai, Y., and Nishizuka, Y. (1977) J. Biol. Chem. 252, 7610-7616), was further degraded by trypsin, whereas the 32-kDa fragment was resistant. Protein kinase activity and the [3H]PDBu binding activity were completely separated upon gel filtration of a solution containing Triton X-100/phosphatidylserine mixed micelles and trypsinized protein kinase C. Pooled fractions of the [3H]PDBu binding activity contained a 32-kDa fragment exclusively. The binding of [3H]PDBu to this fragment was dependent on calcium and phosphatidylserine and was of high affinity (Kd = 2.8 nM) and of essentially identical specificity to that of native protein kinase C. It is concluded that the 32-kDa fragment represents a lipid binding, regulatory domain of protein kinase C.  相似文献   

4.
5.
Four monoclonal antibodies (mAbs), derived from high-responder Biozzi mice immunized with purified protein kinase C, were selected by ELISA and further characterized by immunoblot: 3G12, 5A2, 36G9 are of isotype gamma 1, kappa and 15G4 is of isotype gamma 2b, kappa. Competition analysis between 15G4 and the three other mAbs showed that 15G4 and 3G12 are directed against either the same or overlapping epitope(s). All four mAbs are specific for the bovine gamma isoform of protein kinase C and cross-react with protein kinase C gamma from a variety of animal species. Immunoblot analysis of protein kinase C tryptic fragments revealed that the mAbs recognize the regulatory domain and not the catalytic domain. Two of the mAbs, 36G9 and 5A2, inhibit protein kinase C gamma cofactor-dependent activity (80% and 50% respectively). Consistent with the epitope mapping, none of mAbs inhibit the cofactor-independent catalytic activity of protein kinase C gamma. Competition analysis between these mAbs and phosphatidylserine, 12-O-tetradecanoylphorbol 13-acetate and Ca2+ showed that 36G9 and 5A2 block cofactor binding to protein kinase C gamma. These four mAbs thus interact with at least three distinct epitopes in the regulatory domain of protein kinase C gamma.  相似文献   

6.
Aminoacridines, potent inhibitors of protein kinase C   总被引:4,自引:0,他引:4  
Acridine orange, acridine yellow G, and related compounds potently inhibited protein kinase C (Ca2+/phospholipid-dependent enzyme) activity and phorbol dibutyrate binding. Inhibition was investigated in vitro using Triton X-100 mixed micellar assays (Hannun, Y. A., Loomis, C. R., and Bell, R. M. (1985) J. Biol. Chem. 260, 10039-10043 and Hannun, Y. A., and Bell, R. M. (1986) J. Biol. Chem. 261, 9341-9347). Inhibition by the acridine derivatives was subject to surface dilution; therefore, the relevant concentration unit is mol % rather than the bulk molar concentration. Fifty percent inhibition of protein kinase C activity occurred at concentrations of these compounds comparable to concentrations of sn-1,2-diacylglycerol (DAG) and phosphatidylserine (PS) required for enzyme activation (i.e. 1-6 mol %). The mechanism of inhibition appeared to be complex: both the catalytic and regulatory sites of protein kinase C were affected. Acridine orange was a competitive inhibitor with respect to MgATP when the catalytic fragment of protein kinase C was employed. Inhibition at the active site was overcome by the addition of Triton X-100 micelles or phospholipid vesicles. When the activity of intact protein kinase C was measured, inhibition was noncompetitive with respect to MgATP. Further kinetic analysis suggested a competitive type of inhibition with respect to PS and DAG implying an interaction of acridine compounds with the regulatory lipid cofactors or with the regulatory domain of protein kinase C. This was further supported by demonstrating inhibition of phorbol dibutyrate binding to both protein kinase C and the lipid-binding domain generated by trypsin hydrolysis. Acridine orange and acridine yellow G also inhibited thrombin-induced 40-kDa phosphorylation in human platelets and phorbol dibutyrate binding to platelets. These effects were also subject to surface dilution. These results suggest that acridine derivatives have multiple interactions with protein kinase C with the predominant effect being inhibition of activation within the regulatory domain of the enzyme. Some of the biologic effects of acridine derivatives including anti-tumor action may occur as a consequence of protein kinase C inhibition.  相似文献   

7.
The nonstructural protein 3 (NS3) of hepatitis C virus (HCV) inhibits the nuclear transport and the enzymatic activity of the catalytic subunit of protein kinase A. This inhibition is mediated by an arginine-rich domain localized between amino acids 1487-1500 of the HCV polyprotein. The data presented here indicate that the arginine-rich domain, when embedded in recombinant fragments of NS3, interacts with the catalytic site of protein kinase C (PKC) and inhibits the phosphorylation mediated by this enzyme in vitro and in vivo. Furthermore, a direct binding of PKC to the NS3 fragments leads to an inhibition of the free shuttling of the kinase between the cytoplasm and the particulate fraction. In contrast, a peptide corresponding to the arginine-rich domain (HCV (1487-1500)), despite also being a PKC inhibitor, did not influence the PKC shuttling process and was transported to the particulate fraction by the translocating kinase upon activation with tetradecanoylphorbol-13-acetate. Using the tetradecanoylphorbol-13-acetate -stimulated respiratory burst of NS3-introduced neutrophils as a model system, we could demonstrate that NS3 is able to block PKC-mediated functions within intact cells. Our data support the possibility that NS3 disrupts the PKC-mediated signal transduction.  相似文献   

8.
Monoclonal antibodies (8/1, 10/10, and 25/3) against rat brain type II protein kinase C were used for the immunochemical characterization of this kinase. These antibodies immunoprecipitated the type II protein kinase C in a dose-dependent manner but did neither to the type I nor III isozyme. Immunoblot analysis of the tryptic fragments from protein kinase C revealed that all three antibodies recognized the 27-38-kDa fragments, the phospholipid/phorbol ester-binding domain, but not the 45-48-kDa fragments, the kinase catalytic domain. The immune complexes of the kinase and the antibodies retained 70-80% of the kinase activity which was dependent on Ca2+ and phosphatidylserine and further activated by diacylglycerol or tumor-promoting phorbol ester. With antibody 8/1, the kinetic parameters with respect to Km for ATP and histone and K alpha for phosphatidylserine and phorbol 12,13-dibutyrate were not significantly influenced. However, the antibody causes variable effects on the K alpha for Ca2+ under different assay conditions. When determined in the presence of phosphatidylserine, the K alpha for Ca2+ was reduced by an order of magnitude (37 +/- 8 to 2.0 +/- 1.8 microM); in the presence of phosphatidylserine and phorbol 12,13-dibutyrate, the K alpha for Ca2+ was not significantly altered; and in the presence of phosphatidylserine and dioleoylglycerol, the kinase became an apparently Ca2+-independent enzyme. The effects of antibody 8/1 on the kinetic parameters of the enzyme for phorbol ester binding were different from those for kinase activity. This antibody causes a 20-30% reduction in phorbol ester binding and a 2-fold increase (1.9 +/- 0.2 to 3.9 +/- 0.3 micrograms/ml) in the concentration of phosphatidylserine required for half-maximal binding, but is without significant influence on those parameters for Ca2+ and phorbol 12,13-dibutyrate. The differential effects of antibody 8/1 on kinase activity and phorbol ester binding with respect to the kinetic parameter of phosphatidylserine suggest that the roles of this phospholipid in supporting phorbol ester binding and kinase activation are different. In the presence of the antibody, the autophosphorylations of the phospholipid/phorbol ester-binding domain and the kinase domain were reduced; the reduction was more pronounced for the former than for the latter. These results suggest that the epitope for antibody 8/1 is localized within the phospholipid/phorbol ester-binding domain at the region adjacent to the kinase domain so that the autophosphorylations of both domains are affected.  相似文献   

9.
10.
Receptors for activated protein kinase C (RACKs) have been isolated from the particulate cell fraction of heart and brain. We previously demonstrated that binding of protein kinase C (PKC) to RACKs requires PKC activators and is via a site on PKC that is distinct from the substrate binding site. Here, we examine the possibility that the C2 region in the regulatory domain of PKC is involved in binding of PKC to RACKs. The synaptic vesicle-specific p65 protein contains two regions homologous to the C2 region of PKC. We found that three p65 fragments, containing either one or two of these PKC C2 homologous regions, bound to highly purified RACKs. Binding of the p65 fragments and PKC to RACKs was mutually exclusive; preincubation of RACKs with the p65 fragments inhibited PKC binding, and preincubation of RACKs with PKC inhibited binding of the p65 fragments. Preincubation of the p65 fragments with a peptide resembling the PKC binding site on RACKs also inhibited p65 binding to RACKs, suggesting that PKC and p65 bind to the same or nearby regions on RACKs. Since the only homologous region between PKC and the p65 fragments is the C2 region, these results suggest that the C2 region on PKC contains at least part of the RACK binding site.  相似文献   

11.
Inhibition of protein kinase C by annexin V.   总被引:11,自引:0,他引:11  
Annexin V is a protein of unknown biological function that undergoes Ca(2+)-dependent binding to phospholipids located on the cytosolic face of the plasma membrane. Preliminary results presented herein suggest that a biological function of annexin V is the inhibition of protein kinase C (PKC). In vitro assays showed that annexin V was a specific high-affinity inhibitor of PKC-mediated phosphorylation of annexin I and myosin light chain kinase substrates, with half-maximal inhibition occurring at approximately 0.4 microM. Annexin V did not inhibit epidermal growth factor receptor/kinase phosphorylation of annexin I or cAMP-dependent protein kinase phosphorylation of the Kemptide peptide substrate. Since annexin V purified from both human placenta and recombinant bacteria inhibited protein kinase C activity, it is not likely that the inhibitor activity was associated with a minor contaminant of the preparations. The following results indicated that the mechanism of inhibition did not involve annexin V sequestration of phospholipid that was required for protein kinase C activation: similar inhibition curves were observed as phospholipid concentration was varied from 0 to 800 micrograms/mL; the extent of inhibition was not significantly affected by the order of addition of phospholipid, substrate, or PKC, and the core domain of annexin I was not a high-affinity inhibitor of PKC even though it had similar Ca2+ and phospholipid binding properties as annexin V. These data indirectly indicate that inhibition occurred by direct interaction between annexin V and PKC. Since the concentration of annexin V in many cell types exceeds the amounts required to achieve PKC inhibition in vitro, it is possible that annexin V inhibits PKC in a biologically significant manner in intact cells.  相似文献   

12.
We have examined the phosphorylation of bovine microtubule-associated protein 4 (MAP4), formerly named MAP-U, by protein kinase C (PKC). When MAP4 was incubated with PKC, about 1 mol of phosphate was incorporated/mol of MAP4. Phosphorylation of MAP4 caused a remarkable decrease in the ability of the MAP to stimulate microtubule assembly. MAP4 consists of an amino-terminal projection domain and a carboxyl-terminal microtubule-binding domain. The carboxyl-terminal domain is subdivided into a Pro-rich region and an assembly-promoting (AP) sequence region containing four tandem repeats of AP sequence that is conserved in MAP4, MAP2, and tau [Aizawa et al. (1990) J. Biol. Chem. 265, 13849-13855]. In order to identify the site of MAP4 phosphorylated by PKC, a series of expressed MAP4 fragments was prepared and treated with the kinase. A fragment corresponding to the Pro-rich region (P fragment) was phosphorylated, while fragments corresponding to the projection domain and the AP sequence region were not. In addition, chymotryptic digestion of an authentic MAP4 prephosphorylated by PKC revealed that phosphate was incorporated almost exclusively into a 27-kDa fragment containing the carboxyl-terminal half of the Pro-rich region. We investigated the phosphorylation site in MAP4 using the P fragment and found that Ser815 was phosphorylated almost exclusively. We conclude that the phosphorylation of a single Ser residue in the Pro-rich region negatively regulates the assembly-promoting activity of MAP4.  相似文献   

13.
We have identified two major proteins in human neutrophils that are phosphorylated in vitro by protein kinase C (PKC) as lipocortins III and a fragment of a lipocortin-like 68-kDa protein. In electroporated cells, the 68-kDa protein was phosphorylated during stimulation of the cells with either FMLP or PMA. Lipocortins are of interest because of their Ca2(+)- and phospholipid-dependent actin binding properties and ability to inhibit phospholipase A2. Two crude fractions of enzymes and proteins exposed to [gamma-32]PATP in the presence of Ca2+, Mg2+, phosphatidylserine and 1,2-oleoyl-acetyl-rac-glycerol were analyzed by gel electrophoresis and autoradiography. A number of proteins in a detergent-free fraction, including proteins at 36 and 32 kDa, were phosphorylated in the presence of these cofactors. In contrast, only two major proteins (35 and 32 kDa) were phosphorylated in a detergent-extracted fraction. Phosphorylation of the 36, 35, and 32 kDa proteins required the presence of Ca2+, Mg2+, and phosphatidylserine in our soluble fraction and detergent extract, indicating PKC-dependent phosphorylation. The 32-kDa protein phosphorylated in both the soluble fraction and detergent extract was identified as lipocortin III by immunoprecipitation with a cross-reactive antibody that recognized lipocortin I and comparison of cyanogen bromide (CNBr) cleavage patterns of this protein with a lipocortin III standard. The 68-kDa protein was identified as a lipocortin VI-like protein by immunoprecipitation with anti-calelectrin. Additionally, the CNBr cleavage pattern of the 68-kDa protein was similar to that of the 36-kDa protein phosphorylated in our soluble fraction. Autoradiograms of the 68- and 36-kDa fragments immunoprecipitated from our soluble fraction with anticalelectrin and cleaved with CNBr showed that both of these proteins were phosphorylated in this sample. Because phosphorylation is known to change the functional characteristics of the lipocortins, the potential exists to link PKC and lipocortins in neutrophils to regulation of granulemembrane interactions or mediation of inflammation.  相似文献   

14.
Interactions of types I, II, and III protein kinase C (PKC) with phospholipids were investigated by following the changes in protein kinase activity and phorbol ester binding. The acidic phospholipids such as phosphatidylserine (PS), phosphatidic acid, phosphatidyl-glycerol, and cardiolipin, which are activators of PKC in the assay of protein phosphorylation, could differentially inactivate PKC I, II, and III during preincubation in the absence of divalent cation. The phospholipid-induced inactivation of PKC was concentration and time dependent and only affected the kinase activity without influencing phorbol ester binding. PKC I was the most susceptible to the phospholipid-induced inactivation, and PKC III was the least. The IC50 values of PS for PKC I, II, and III were 5, 45, and greater than 120 microM, respectively. Addition of divalent cation such as Ca2+ or Mg2+ suppressed the phospholipid-induced inactivation of PKC. In the absence of divalent cation, PKC I, II, and III all formed complexes with PS vesicles, although to a slightly different degree, as analyzed by molecule sieve chromatography. [3H]Phorbol 12,13-dibutyrate binding for PKC I, II, and III was recovered after chromatography; however, the kinase activities of all these enzymes were greatly reduced. In the presence of Ca2+, all three PKCs formed complexes with PS vesicles, and both the kinase and phorbol ester-binding activities of PKC II and III were recovered following chromatography. Under the same conditions, the phorbol ester-binding activity of PKC I was also recovered, but the kinase activity was not. The phospholipid-induced inactivation of PKC apparently results from a direct interaction of phospholipid with the catalytic domain of PKC; this interaction can be suppressed by divalent cations. In the presence of divalent cations, PS interacted preferentially with the regulatory domain of PKC and resulted in the activation of the kinase.  相似文献   

15.
Chelerythrine is a potent and specific inhibitor of protein kinase C   总被引:56,自引:0,他引:56  
The benzophenanthridine alkaloid chelerythrine is a potent, selective antagonist of the Ca++/phospholopid-dependent protein kinase (Protein kinase C: PKC) from the rat brain. Half-maximal inhibition of the kinase occurs at 0.66 microM. Chelerythrine interacted with the catalytic domain of PKC, was a competitive inhibitor with respect to the phosphate acceptor (histone IIIS) (Ki = 0.7 microM) and a non-competitive inhibitor with respect to ATP. This effect was further evidenced by the fact that chelerythrine inhibited native PKC and its catalytic fragment identically and did not affect [3H]- phorbol 12,13 dibutyrate binding to PKC. Chelerythrine selectively inhibited PKC compared to tyrosine protein kinase, cAMP-dependent protein kinase and calcium/calmodulin-dependent protein kinase. The potent antitumoral activity of celerythrine measured in vitro might be due at least in part to inhibition of PKC and thus suggests that PKC may be a model for rational design of antitumor drugs.  相似文献   

16.
Limited tryptic digestion of protein kinase C purified from mouse brain generated a 36-kDa fragment which no longer required Ca2+ and phospholipid for activity or bound phorbol ester. Under appropriate conditions, the isolated fragment was stable for several months at 4 degrees C or upon freezing and storage at -70 degrees C. Kinetic characteristics of the fragment were similar to those for the intact protein kinase. Although the fragment did not require phospholipid for activity, anionic phospholipids affected the extent of its activity in a pH-, substrate-, and substrate concentration-dependent manner. This effect appeared to be due to complex formation between the phospholipid and substrate. The catalytic fragment thus permits detection of a second point of interaction of phospholipid with the protein kinase C system in addition to the already described phospholipid regulatory domain.  相似文献   

17.
In fibroblasts, the protein kinase C (PKC) activator phorbol 12-myristate 13-acetate (PMA) stimulates phospholipase D (PLD)-mediated hydrolysis of both phosphatidylcholine (PtdCho) and phosphatidylethanolamine (PtdEtn) by PKC-alpha-mediated nonphosphorylating and phosphorylating mechanisms. Here we have used NIH 3T3 fibroblasts overexpressing holo PKC-epsilon and its regulatory, catalytic, and zinc finger domain fragments to determine if this isozyme also regulates PLD activity. Overexpression of holo PKC-epsilon inhibited the stimulatory effects of PMA (5-100 nM) on both PtdCho and PtdEtn hydrolysis. Overexpression of PKC-epsilon also was found to inhibit platelet-derived growth factor-induced PLD activity. Expression of the catalytic unit of PKC-epsilon had no effect on PMA-induced PLD activity. In contrast, expression of both the regulatory domain fragment and the zinc finger domain of PKC-epsilon resulted in significant inhibition of PMA-stimulated PtdCho and PtdEtn hydrolysis. Interestingly, although PKC-alpha also mediates the stimulatory effect of PMA on the synthesis of PtdCho by a phosphorylation mechanism, overexpression of holo PKC-epsilon or its regulatory domain fragments did not affect PMA-induced PtdCho synthesis. These results indicate that the PKC-epsilon system can act as a negative regulator of PLD activity and that this inhibition is mediated by its regulatory domain.  相似文献   

18.
The binding of pyridoxal analogues to the structural domains of pyridoxal kinase was studied by fluorescence spectroscopy and chromatographic techniques. Two fragments of 24 and 16 kDa, arising from limited proteolysis of the native enzyme, were separated by ion-exchange chromatography and used for binding studies with pyridoxal oxime. Fluorometric titrations yielded dissociation constants of 6 and 12.4 MicroM for pyridoxal oxime bound to the native enzyme and 24-kDa fragment, respectively. 4-(4-Azido-2-nitrophenyl)-pyridoxamine, a new photolabeling reagent, binds irreversibly to the kinase with concomitant loss of catalytic activity. The modified kinase (2.1 mol label/mol dimer) yields two fragments upon limited proteolysis with chymotrypsin. The two fragments were separated by reverse-phase HPLC and SDS/polyacrylamide gel electrophoresis. Radiolabeled ligand was detected only in the 24-kDa fragment. It is postulated that the pyridoxal binding site is located in the 24-kDa structural domain.  相似文献   

19.
Protein kinase C (PKC) has been implicated in a variety of cellular responses such as proliferation, differentiation, and secretion. We assessed the role of PKC in the mitogenic effects of gastrin-releasing peptide (in a small cell lung cancer (SCLC) cell line. Using antisera that specifically recognize the PKC isoforms alpha, beta, gamma, delta, and epsilon, we determined that PKC epsilon is the major isoform in the SCLC cell line NCI-N417, followed by PKC alpha and delta. In addition to the 90-kDa PKC epsilon, our anti-PKC epsilon antiserum specifically detected a 40-kDa immunoreactive protein. Treatment of the cells with either 20 nM phorbol myristate acetate or 50 nM GRP enhanced significantly the level of the 40-kDa protein in a time-dependent (1-8 h), cycloheximide-sensitive fashion. Subcellular fractionation revealed that 90% of PKC epsilon was in particulate form, while the 40-kDa immunoreactive protein was cytosolic. To test the hypothesis that the 40-kDa soluble protein represented a catalytically independent PKC epsilon fragment, cytosolic extracts were assayed for kinase activity. 45-50% of the activity was apparent in the absence of the PKC activators phosphatidylserine and diacylglycerol. This effector-independent kinase activity was further purified by affinity chromatography using a synthetic peptide corresponding to the pseudosubstrate region of PKC epsilon (ERMRPRKRQGAVRRRV) coupled to Sepharose. The partially purified protein, recognized by the anti-PKC epsilon antiserum, exhibited histone kinase activity with kinetics similar to those of the tryptically generated catalytic fragment of brain PKC epsilon. This activity was inhibited by staurosporine (IC50 = 1 x 10(-8) M) and by the pseudosubstrate inhibitor peptide (IC50 = 7.7 x 10(-8) M). The SCLC kinase and the brain PKC epsilon catalytic fragment were similar as indicated by the relative sizes of the PKC epsilon immunoreactive peptides generated with protease V8 from Staphylococcus aureus (Mr approximately 37,000, 34,000, 28,000, 26,000, and 25,000). Taken together, we conclude that a variant SCLC cell line expresses a constitutively active catalytic fragment of PKC epsilon. Regulation by 12-O-tetradecanoyl-13-acetate or GRP via de novo protein synthesis suggests a novel mechanism of control of PKC diversity with implications for small cell lung cancer and possibly other malignancies.  相似文献   

20.
The oxidant mitogen/tumor promoter, periodate, was used to selectively modify either the regulatory domain or the catalytic domain of protein kinase C (PKC) to induce oxidative activation or inactivation of PKC, respectively. Periodate, at micromolar concentrations, modified the regulatory domain of PKC as determined by the loss of ability to stimulate kinase activity by Ca2+/phospholipid, and also by the loss of phorbol ester binding. This modification resulted in an increase in Ca2+/phospholipid-independent kinase activity (oxidative activation). However, at higher concentrations (greater than 100 microM) periodate also modified the catalytic domain, resulting in complete inactivation of PKC. The oxidative modification induced by low periodate concentrations (less than 0.5 mM) was completely reversed by a brief treatment with 2 mM dithiothreitol. In this aspect, the modification induced by periodate was different from that of the previously reported irreversible modification of PKC induced by H2O2. However, the inactivation of PKC induced by periodate at concentrations greater than 1 mM was not reversed by dithiothreitol. Among the phospholipids and ligands of the regulatory domain tested, only phosphatidylserine protected the regulatory domain from oxidative modification. In the presence of phosphatidylserine, the catalytic site was selectively modified by periodate, resulting in formation of a form of PKC that exhibited phorbol ester binding but not kinase activity. Both reversible and irreversible oxidative activation and inactivation of PKC also were observed in intact cells treated with periodate. Taken together these results suggest that periodate, by virtue of having a tetrahedral structure, binds to the phosphate-binding regions present within the phosphatidylserine-binding site of the regulatory domain and the ATP-binding site of the catalytic domain, and modifies the vicinal thiols present within these sites. This results in the formation of intramolecular disulfide bridge(s) within the regulatory domain or catalytic domain leading to either reversible activation or inactivation of PKC, respectively. Thus, oxidant mitogen/tumor promoters such as periodate may be able to bypass normal transmembrane signalling systems to directly activate pathways involved in cellular regulation.  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