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.
2.
Melanogenesis is regulated by a variety of environmental and hormonal factors. In this study, we showed that protein kinase C (PKC) plays a major role in regulating melanogenesis in B16 mouse melanoma cells. Chronic treatment of B16 cells with phorbol dibutyrate resulted in a concentration-dependent loss of density-dependent induction of tyrosinase activity, which correlated positively with a concentration-dependent loss of PKC enzyme activity. In contrast, B16 clones overexpressing PKCα had increased tyrosinase activity. Different phorbol derivatives inhibited tyrosinase activity and depleted cellular PKCα in a manner that reflected their reported tumor-promoting activity. Western blotting analysis showed that phorbol dibutyrate decreased the amount of the brown locus gene product (TRP-1) by 50% and lowered the amount of the albino locus gene product (tyrosinase) to undetectable levels. None of the phorbol derivatives affected the level of the slaty locus protein (TRP-2). The decrease in tyrosinase and TRP-1 protein levels was found to be due to a decrease in the mRNA encoded by these genes. In addition to inhibiting the density-dependent increase in tyrosinase activity, phorbol dibutyrate inhibited some, but not all, of the 8-bromocyclic AMP-induced increase in tyrosinase activity. This was accompanied by a decrease in the amount of tyrosinase protein induced by 8-bromocyclic AMP. Although 8-bromocyclic AMP did not change the level of TRP-1, it did reverse the decrease in the amount of this protein induced by phorbol dibutyrate. The amount of TRP-2 was not altered by any of these agents. These data suggest that PKC regulates melanogenesis primarily by controlling the constitutive expression of tyrosinase and, to a lesser extent, TRP-1. © 1996 Wiley-Liss, Inc.  相似文献   

3.
4.
Retinoic acid receptor (RAR) α and γ mRNAs were constitutively expressed in B16 melanoma cells with or without retinoic acid (RA) treatment. RARβ mRNA, however, was significantly expressed only after exposure to RA. Induction of RARβ by RA occurred within 1 h and was not inhibited by cycloheximide (i.e., did not require new protein synthesis). All three RAR mRNA levels were dramatically decreased with 8-bromo-cyclic AMP treatment and could not be rescued by addition of RA. Analysis of RARγ revealed that this decrease occurred within 1 h of exposure to 8-bromo-cyclic AMP and was not blocked by simultaneous treatment with cycloheximide. The stability of RARγ mRNA was not altered by cyclic AMP treatment. Nuclear extracts from 8-bromo-cyclic AMP-treated cells showed a large decrease in protein binding to a retinoic acid response element (RARE) oligonucleotide compared to control cells. This correlated with a marked reduction of RA-stimulated RARE-reporter gene activity in transfected cells which were treated with cyclic AMP. Pretreatment of B16 cells with cyclic AMP prior to RA addition dramatically reduced induction of PKCα, an early marker of RA-induced cell differentiation. Thus, cyclic AMP can antagonize the action of RA most likely via its ability to inhibit RAR expression. © 1996 Wiley-Liss, Inc.  相似文献   

5.
We previously reported that retinoic acid (RA) augmented mouse (BALB/c) lymphokine (interleukin-2)-activated killer (LAK) cell activity in a dose and time dependent manner. As evidence available has suggested the role of protein kinase C (PKC) in the regulation of cell mediated cytotoxicity, the present work was to investigate whether or not PKC may mediate the enhancement of LAK cell activity by RA. Accompanied with an augmented LAK cell activity, RA increased total PKC enzyme activity, [3H]phorbol 12,13-dibutyrate binding activity, and the amount of immunoreactive PKC. A prolonged treatment (18 h) of LAK cells with 12-O-tetradecanoylphorbol-13-acetate resulted in the loss of both PKC and LAK cell activity. PKC inhibitors, 1-(5-isoquinolinesulfonyl)-2-methyl-piperazine dihydrochloride and staurosporine, also drastically reduced LAK cell activity. Although most of the total PKC activity (97%) was detected in the cytosol fraction, the increase in PKC activity was attributed to an increased enzyme activity in both cytosol and membrane fractions, and shown to be RA dose-dependent. Kinetics study revealed that the increase in PKC was a time-dependent process and the enhancement was detectable as early as 8 h after the addition of RA to LAK cell culture. By immunoblotting, the cytosol PKC of LAK cells was shown to contain alpha and beta isoforms, but not gamma. RA further increased the expression of PKC alpha. The enhanced expression of alpha isozyme of PKC by RA was also in a dose and time dependent manner. Taken together, these results indicate that the mechanism of the augmentation of LAK cell activity by RA may in part result from the increase in PKC, especially PKC alpha isozyme.  相似文献   

6.
Vitamin A inhibits growth and increases the activity of cAMP-dependent protein kinase in B16 mouse melanoma cells. In this report we show that retinoic acid (RA) treatment of intact cells alters their subsequent in vitro protein phosphorylation, but we could not demonstrate any changes in in vivo protein phosphorylation. A 48-h treatment with RA results in a concentration-dependent decrease of protein phosphorylation of a 95K molecular weight (MW) protein in both supernatant and particulate fractions. The phosphorylation of this protein does not appear to be regulated by cAMP. Proteins at 92K and 82K MW in the supernatant fraction are increased in phosphorylation. The former (but not the latter) is regulated by cAMP. In the particulate fraction a variety of proteins 12K-68K MW are increased in phosphorylation, as the cells are treated with increasing amounts of RA. The phosphorylation of most of these proteins is regulated by cAMP. Another inhibitor of B16 cell growth, melanocyte-stimulating hormone (MSH) also alters protein phosphorylation. At short incubation periods (1 h), this hormone stimulates phosphorylation of a number of proteins (17-40K MW), while in longer incubation periods (48 h) phosphorylation is inhibited. All of these phosphorylations appear to be regulated by cAMP. We attempted to repeat these observations using intact-cell phosphorylation with 32PO4. In two experiments we saw small changes in the phosphorylation of proteins. In most experiments, however, we could find no change in the phosphoproteins. Further experiments have led us to question the in vivo phosphorylation, since treatment of the cells with MSH, cholera toxin, or db-cAMP also did not affect intact-cell protein phosphorylation. We have previously documented that under these latter conditions cAMP levels are greatly elevated and cAMP-dependent protein kinase is activated. The in vitro phosphorylation results suggests that in RA-treated cells, kinase activities and/or protein substrate levels are changing. However, the physiological significance of the particular MW phosphoproteins changes we have described must await resolution of the in vivo phosphorylation data.  相似文献   

7.
8.
Mitogen-activated protein kinase (MAPK) and protein phosphatase 2A (PP2A) regulate oocyte meiosis, yet little is known regarding their mechanisms of action. This study addressed the functional importance of active MAPK and PP2A in regulating oocyte meiosis. Experiments were conducted to identify MAPK activation, PP2A activity, intracellular enzyme trafficking, and ultrastructural associations during meiosis. Questions of requisite kinase and/or phosphatase activity and chromatin condensation, microtubule polymerization, and spindle formation were addressed. At the protein level, MAPK and PP2A were present in constant amounts throughout the first meiotic division. Both MAPK and PP2A were activated following germinal vesicle breakdown (GVBD) in conjunction with metaphase I development. Immunocytochemical studies confirmed the absence of active MAPK in germinal vesicle-intact (GVI) and GVBD oocytes. At metaphase I and during the metaphase I/metaphase II transition, activated MAPK colocalized with microtubules, poles, and plates of meiotic spindles. Protein phosphatase 2A was dispersed evenly throughout the GVI oocyte cytoplasm. Throughout the metaphase I/metaphase II transition, PP2A colocalized with microtubules of meiotic spindles. Both active MAPK and PP2A associated with in vitro-polymerized microtubules, suggesting that active MAPK and PP2A locally regulate spindle formation. Inhibition of MAPK activation resulted in compromised microtubule polymerization, no spindle formation, and loosely condensed chromosomes. Treatment with okadaic acid (OA) or calyculin-A (CL-A), which inhibits oocyte cytoplasmic PP2A, caused an absence of microtubule polymerization and spindles, even though MAPK activity was increased under these treatment conditions. Thus, active MAPK is required, but is not sufficient, for normal meiotic spindle formation and chromosome condensation. In addition, the oocyte OA/CL-A-sensitive PP, presumably PP2A, is essential for microtubule polymerization and meiotic spindle formation.  相似文献   

9.
Glial cell line‐derived neurotrophic factor (GDNF) and retinoic acid (RA) are two molecules crucial for the regulation of the spermatogonial compartment of the testis. During the cycle of the seminiferous epithelium, their relative concentration oscillates with lower GDNF levels in stages where RA levels are high. It has been recently shown that RA negatively regulates Gdnf expression but the mechanisms behind are so far unknown. Here, we show that RA directly downregulates Gdnf mRNA levels in primary murine Sertoli cells through binding of RARα to a novel DR5‐RARE on Gdnf promoter. Pharmacological inhibition and chromatin immunoprecipitation–quantitative polymerase chain reaction analysis suggested that the underlying mechanism involved histone deacetylase activity and epigenetic repression of Gdnf promoter upon RA treatment.  相似文献   

10.
Regulation of VL30 gene expression by activators of protein kinase C   总被引:9,自引:0,他引:9  
The mouse genome contains a retrovirus-like sequence, designated VL30, which is expressed at high levels in transformed cells and which can be induced by exogenously supplied epidermal growth factor (EGF). Binding of EGF to the EGF receptor produces changes in intracellular calcium levels and phospholipase activity which indirectly lead to activation of protein kinase C. We treated AKR-2B cells, Swiss 3T3 cells, and the 3T3 variants NR6 (EGF receptorless) and TNR9 (phorbol ester nonresponsive) with various phorbol ester tumor promoters and with the synthetic diacylglycerol sn-1,2-dioctanoylglycerol. Tumor-promoting phorbol esters (e.g. 12-O-tetradecanoyl phorbol acetate (TPA] increased the level of VL30 expression. Stimulation with either TPA or EGF produced a similar time course of VL30 expression. TPA induced VL30 expression in the EGF-receptorless NR6 cell line, indicating that neither EGF ligand-receptor binding nor phosphorylation of the EGF receptor was required for induction of VL30 expression. Protein synthesis was not required for the TPA-mediated increase in VL30 expression, as pretreatment with cycloheximide did not block or reduce the TPA effect. VL30 expression was also stimulated by treatment with sn-1,2-dioctanoylglycerol, an analog of a probable endogenous activator of protein kinase C. These results suggest that activation of protein kinase C plays a direct role in regulating VL30 expression.  相似文献   

11.
12.
13.
Differentiation of B16 mouse melanoma cells induced by retinoic acid (RA) is preceded by a large increase in protein kinase C alpha (PKC alpha) mRNA and protein. To determine the role of PKC alpha in the differentiation program, we stably transfected B16-F1 cells with a plasmid containing the full length PKC alpha cDNA driven by an SV40 promoter. Two out of thirty-two colonies screened were determined to overexpress PKC by 2-4-fold according to Western blot analysis and PKC enzyme activity. When compared to control cells (wild-type cells and cells transfected only with the neomycin resistance gene), PKC alpha overexpressing clones displayed longer doubling times, diminished anchorage-independent growth, and increased melanin production. RA treatment of control cells mimicked these phenotypic characteristics. When injected subcutaneously into syngeneic mice, PKC alpha overexpressing clones produced smaller tumors and had longer latencies than control cells. These findings, combined with the fact that phorbol esters down-regulate PKC and antagonize RA action suggest that PKC alpha plays a key role in the RA-induced melanoma differentiation.  相似文献   

14.
15.
16.

Background  

We have previously shown that ultraviolet-A (UVA) radiation enhances metastatic lung colonization capacity of B16-F1 melanoma cells. The aim of this study was to examine changes in expression profile of genes in mouse melanoma B16-F1 cells exposed to UVA radiation.  相似文献   

17.
Resident mouse peritoneal macrophages synthesized and released prostaglandins (PGs) when challenged with 12-O-tetradecanoylphorbol 13-acetate (TPA) or 1,2-dioctanoyl-sn-glycerol (DiC8). Both stimuli were found to activate Ca2+/phospholipid-dependent protein kinase C (PKC). 1-(5-Isoquinolinesulphonyl)-2-methylpiperazine ('H-7') and D-sphingosine, known to inhibit PKC by different mechanisms, were able to decrease the PKC activity of macrophages in a dose-dependent manner. Addition of either PKC inhibitor decreased PG synthesis and also the release of arachidonic acid (AA) from phospholipids induced by TPA or DiC8. Simultaneously TPA or DiC8 also decreased incorporation of free AA into membrane phospholipids of macrophages. AA incorporation could be restored, however, by pretreatment with the PKC inhibitors. Our results demonstrate an involvement of PKC in the regulation of PG synthesis in mouse peritoneal macrophages and provide further evidence that reacylation of released fatty acids may be an important regulatory step.  相似文献   

18.
Our previous studies showed that docetaxel-induced apoptosis of human melanoma cells was dependent on the activation of the c-jun NH(2)-terminal kinase (JNK) signaling pathway but was inhibited by the extracellular signal-regulated kinase (ERK)-1/2 pathway. However, the mechanisms by which these pathways were modulated by docetaxel were not clear. We report here that docetaxel induces activation of protein kinase C (PKC) signaling differentially through PKCepsilon and PKCdelta isoforms. Activation of PKCepsilon was most marked in docetaxel-resistant cells and paralleled the activation of the ERK1/2 pathway. Inhibition of PKCepsilon by small interfering RNA molecules resulted in down-regulation of phosphorylated ERK1/2 and sensitization of cells to docetaxel-induced apoptosis. Experiments also showed that beta-tubulin class III, a molecular target of docetaxel, coimmunoprecipitated with PKCepsilon and colocalized in confocal microscopic studies. In contrast to PKCepsilon, high levels of activated PKCdelta were associated with activation of the JNK pathway and sensitivity to docetaxel. Activation of PKCdelta seemed to be upstream of JNK because inhibition of PKCdelta by small interfering RNA abrogated activation of the JNK pathway. Although PKCdelta could be activated in resistant cells, downstream activation of JNK and c-Jun did not occur. In summary, these results suggest that the outcome of docetaxel-induced apoptotic events in human melanoma cells depends on their PKC isoform content and signaling responses. PKCepsilon was associated with prosurvival signaling through ERK, whereas PKCdelta was associated with proapoptotic responses through JNK activation.  相似文献   

19.
The glycopeptides produced by B16 mouse melanoma cells grown in the presence of [3H]glucosamine were isolated and fractionated into two classes (I and II) with cetyl pyridinium chloride. The class I glycopeptides were of higher molecular weight and of higher negative charge (sialic acid content) than those in class II. Class I glycopeptides contained N-acetyl neuraminic acid, galactose and N-acetylgalactosamine and on treatment with alkaline-borohydride were degraded to apparently tri- and tetrasaccharides. The presence of this mucin-type glycoprotein on the cell surface was detected by mild trypsinization of intact cells.  相似文献   

20.
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