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.
Vitamin D metabolites 1alpha,25(OH)(2)D(3) and 24R,25(OH)(2)D(3) regulate endochondral ossification in a cell maturation-dependent manner via membrane-mediated mechanisms. 24R,25(OH)(2)D(3) stimulates PKC activity in chondrocytes from the growth plate resting zone, whereas 1alpha,25(OH)(2)D(3) stimulates PKC in growth zone chondrocytes. We used the rat costochondral growth plate cartilage cell model to study how these responses are differentially regulated. 1alpha,25(OH)(2)D(3) acts on PKC, MAP kinase, and downstream physiological responses via phosphatidylinositol-specific PLC-beta; 24R,25(OH)(2)D(3) acts via PLD. In both cases, diacylglycerol (DAG) is increased, activating PKC. Both cell types possess membrane and nuclear receptors for 1alpha,25(OH)(2)D(3), but the mechanisms that render the 1alpha,25(OH)(2)D(3) pathway silent in resting zone cells or the 24R,25(OH)(2)D(3) pathway silent in growth zone cells are unclear. PLA(2) is pivotal in this process. 1alpha,25(OH)(2)D(3) stimulates PLA(2) activity in growth zone cells and 24R,25(OH)(2)D(3) inhibits PLA(2) activity in resting zone cells. Both processes result in PKC activation. To understand how negative regulation of PLA(2) results in increased PKC activity in resting zone cells, we used PLA(2) activating peptide to stimulate PLA(2) activity and examined cell response. PLAP is not expressed in resting zone cells in vivo, supporting the hypothesis that PLA(2) activation is inhibitory to 24R,25(OH)(2)D(3) action in these cells.  相似文献   

3.
1alpha,25(OH)(2)D(3) regulates rat growth plate chondrocytes via nuclear vitamin D receptor (1,25-nVDR) and membrane VDR (1,25-mVDR) mechanisms. To assess the relationship between the receptors, we examined the membrane response to 1alpha,25(OH)(2)D(3) in costochondral cartilage cells from wild type VDR(+/+) and VDR(-/-) mice, the latter lacking the 1,25-nVDR and exhibiting type II rickets and alopecia. Methods were developed for isolation and culture of cells from the resting zone (RC) and growth zone (GC, prehypertrophic and upper hypertrophic zones) of the costochondral cartilages from wild type and homozygous knockout mice. 1alpha,25(OH)(2)D(3) had no effect on [(3)H]-thymidine incorporation in VDR(-/-) GC cells, but it increased [(3)H]-thymidine incorporation in VDR(+/+) cells. Proteoglycan production was increased in cultures of both VDR(-/-) and VDR(+/+) cells, based on [(35)S]-sulfate incorporation. These effects were partially blocked by chelerythrine, which is a specific inhibitor of protein kinase C (PKC), indicating that PKC-signaling was involved. 1alpha,25(OH)(2)D(3) caused a 10-fold increase in PKC specific activity in VDR(-/-), and VDR(+/+) GC cells as early as 1 min, supporting this hypothesis. In contrast, 1alpha,25(OH)(2)D(3) had no effect on PKC activity in RC cells isolated from VDR(-/-) or VDR(+/+) mice and neither 1beta,25(OH)(2)D(3) nor 24R,25(OH)(2)D(3) affected PKC in GC cells from these mice. Phospholipase C (PLC) activity was also increased within 1 min in GC chondrocyte cultures treated with 1alpha,25(OH)(2)D(3). As noted previously for rat growth plate chondrocytes, 1alpha,25(OH)(2)D(3) mediated its increases in PKC and PLC activities in the VDR(-/-) GC cells through activation of phospholipase A(2) (PLA(2)). These responses to 1alpha,25(OH)(2)D(3) were blocked by antibodies to 1,25-MARRS, which is a [(3)H]-1,25(OH)(2)D(3) binding protein identified in chick enterocytes. 24R,25(OH)(2)D(3) regulated PKC in VDR(-/-) and VDR(+/+) RC cells. Wild type RC cells responded to 24R,25(OH)(2)D(3) with an increase in PKC, whereas treatment of RC cells from mice lacking a functional 1,25-nVDR caused a time-dependent decrease in PKC between 6 and 9 min. 24R,25(OH)(2)D(3) dependent PKC was mediated by phospholipase D, but not by PLC, as noted previously for rat RC cells treated with 24R,25(OH)(2)D(3). These results provide definitive evidence that there are two distinct receptors to 1alpha,25(OH)(2)D(3). 1alpha,25(OH)(2)D(3)-dependent regulation of DNA synthesis in GC cells requires the 1,25-nVDR, although other physiological responses to the vitamin D metabolite, such as proteoglycan sulfation, involve regulation via the 1,25-mVDR.  相似文献   

4.
1alpha,25(OH)(2)D(3) activates protein kinase C (PKC) in rat growth plate chondrocytes via mechanisms involving phosphatidylinositol-specific phospholipase C (PI-PLC) and phospholipase A(2) (PLA(2)). The purpose of this study was to determine if 1alpha,25(OH)(2)D(3) activates PI-PLC directly or through a PLA(2)-dependent mechanism. We determined which PLC isoforms are present in the growth plate chondrocytes, and determined which isoform(s) of PLC is(are) regulated by 1alpha,25(OH)(2)D(3). Inhibitors and activators of PLA(2) were used to assess the inter-relationship between these two phospholipid-signaling pathways. PI-PLC activity in lysates of prehypertrophic and upper hypertrophic zone (growth zone) cells that were incubated with 1alpha,25(OH)(2)D(3), was increased within 30s with peak activity at 1-3 min. PI-PLC activity in resting zone cells was unaffected by 1alpha,25(OH)(2)D(3). 1beta,25(OH)(2)D(3), 24R,25(OH)(2)D(3), actinomycin D and cycloheximide had no effect on PLC in lysates of growth zone cells. Thus, 1alpha,25(OH)(2)D(3) regulation of PI-PLC enzyme activity is stereospecific, cell maturation-dependent, and nongenomic. PLA(2)-activation (mastoparan or melittin) increased PI-PLC activity to the same extent as 1alpha,25(OH)(2)D(3); PLA(2)-inhibition (quinacrine, oleyloxyethylphosphorylcholine (OEPC), or AACOCF(3)) reduced the effect of 1alpha,25(OH)(2)D(3). Neither arachidonic acid (AA) nor its metabolites affected PI-PLC. In contrast, lysophosphatidylcholine (LPC) and lysophosphatidylethanolamine (LPE) activated PI-PLC (LPE>LPC). 1alpha,25(OH)(2)D(3) stimulated PI-PLC and PKC activities via Gq; GDPbetaS inhibited activity, but pertussis toxin did not. RT-PCR showed that the cells express PLC-beta1a, PLC-beta1b, PLC-beta3 and PLC-gamma1 mRNA. Antibodies to PLC-beta1 and PLC-beta3 blocked the 1alpha,25(OH)(2)D(3) effect; antibodies to PLC-delta and PLC-gamma did not. Thus, 1alpha,25(OH)(2)D(3) regulates PLC-beta through PLA(2)-dependent production of lysophospholipid.  相似文献   

5.
Phospholipase A2 (PLA2) is pivotal in the rapid membrane-mediated actions of 1,25-dihydroxyvitamin D3 [1alpha,25(OH)2D3]. Microarray analysis indicated that PLA2 activating protein (PLAA) mRNA is upregulated 6-fold before rat growth plate cells exhibit 1alpha,25(OH)2D3-dependent protein kinase C (PKC) increases, suggesting that it plays an important role in 1alpha,25(OH)2D3's mechanism of action. PLAA mRNA was confirmed in 1alpha,25(OH)2D3-responsive growth zone (prehypertrophic and upper hypertrophic cell zones) chondrocytes by RT-PCR and Northern blot in vitro and by in situ hybridization in vivo. PLAA protein was shown by Western blot and immunohistochemistry. PLAAs role in 1alpha,25(OH)2D3 signaling was evaluated in growth zone cell cultures using PLAA peptide. Arachidonic acid release was increased as was PLA2-specific activity in plasma membranes and matrix vesicles. PKCalpha, but not PKCbeta, PKCepsilon, or PKCzeta, was increased. PLAAs effect was comparable to that of 1alpha,25(OH)2D3 and was additive with 1alpha,25(OH)2D3. PLA2 inhibitors quinacrine and AACOCF3, and cyclooxygenase inhibitor indomethacin blocked the effect of PLAA peptide on PKC, indicating arachidonic acid and its metabolites were involved. This was confirmed using exogenous arachidonic acid. Prostaglandin acted via EP1 based on inhibition by SC19220 and not via EP2 since AH6809 had no effect. Like 1alpha,25(OH)2D3, PLAA peptide also increased activity of phospholipase C-specific activity via beta-1 and beta-3 isoforms, but not delta-1 or gamma-1; the effect of PLAA was via lysophospholipid but not via arachidonic acid. PLAA peptide decreased [3H]-thymidine incorporation to 50% of the decrease caused by 1alpha,25(OH)2D3. In contrast, PLAA peptide increased alkaline phosphatase-specific activity and proteoglycan production in a manner similar to 1alpha,25(OH)2D3. This indicates that PLAA is a specific activator of PLA2 in growth plate chondrocytes, and suggests that it mediates the membrane effect of 1alpha,25(OH)2D3, thereby modulating physiological response.  相似文献   

6.
Growth plate chondrocytes produce proteoglycan-rich type II collagen extracellular matrix (ECM). During cell maturation and hypertrophy, ECM is reorganized via a process regulated by 1alpha,25(OH)(2)D(3) and involving matrix metalloproteinases (MMPs), including MMP-3 and MMP-2. 1alpha,25(OH)(2)D(3) regulates MMP incorporation into matrix vesicles (MVs), where they are stored until released. Like plasma membranes (PM), MVs contain the 1alpha,25(OH)(2)D(3)-binding protein ERp60, phospholipase A(2) (PLA(2)), and caveolin-1, but appear to lack nuclear Vitamin D receptors (VDRs). Chondrocytes produce 1alpha,25(OH)(2)D(3) (10(-8)M), which binds ERp60, activating PLA(2), and resulting lysophospholipids lead to MV membrane disorganization, releasing active MMPs. MV MMP-3 activates TGF-beta1 stored in the ECM as large latent TGF-beta1 complexes, consisting of latent TGF-beta1 binding protein, latency associated peptide, and latent TGF-beta1. Others have shown that MMP-2 specifically activates TGF-beta2. TGF-beta1 regulates 1alpha,25(OH)(2)D(3)-production, providing a mechanism for local control of growth factor activation. 1alpha,25(OH)(2)D(3) activates PKCalpha in the PM via ERp60-signaling through PLA(2), lysophospholipid production, and PLCbeta. It also regulates distribution of phospholipids and PKC isoforms between MVs and PMs, enriching the MVs in PKCzeta. Direct activation of MMP-3 in MVs requires ERp60. However, when MVs are treated with 1alpha,25(OH)(2)D(3), PKCzeta activity is decreased and PKCalpha is unaffected, suggesting a more complex feedback mechanism, potentially involving MV lipid signaling.  相似文献   

7.
1alpha,25-(OH)(2)D(3) regulates protein kinase C (PKC) activity in growth zone chondrocytes by stimulating increased phosphatidylinositol-specific phospholipase C (PI-PLC) activity and subsequent production of diacylglycerol (DAG). In contrast, 24R,25-(OH)(2)D(3) regulates PKC activity in resting zone (RC) cells, but PLC does not appear to be involved, suggesting that phospholipase D (PLD) may play a role in DAG production. In the present study, we examined the role of PLD in the physiological response of RC cells to 24R,25-(OH)(2)D(3) and determined the role of phospholipases D, C, and A(2) as well as G-proteins in mediating the effects of vitamin D(3) metabolites on PKC activity in RC and GC cells. Inhibition of PLD with wortmannin or EDS caused a dose-dependent inhibition of basal [3H]-thymidine incorporation by RC cells and further increased the inhibitory effect of 24R,25-(OH)(2)D(3). Wortmannin also inhibited basal alkaline phosphatase activity and [35]-sulfate incorporation and decreased the stimulatory effect of 24R,25-(OH)(2)D(3). This inhibitory effect of wortmannin was not seen in cultures treated with the PI-3-kinase inhibitor LY294002, verifying that wortmannin affected PLD. Wortmannin also inhibited basal PKC activity and partially blocked the stimulatory effect of 24R,25-(OH)(2)D(3) on this enzyme activity. Neither inhibition of PI-PLC with U73122, nor PC-PLC with D609, modulated PKC activity. Wortmannin had no effect on basal PLD in GC cells, nor on 1alpha,25-(OH)(2)D(3)-dependent PKC. Inhibition of PI-PLC blocked the 1alpha,25-(OH)(2)D(3)-dependent increase in PKC activity but inhibition of PC-PLC had no effect. Activation of PLA(2) with melittin inhibited basal and 24R,25-(OH)(2)D(3)-stimulated PKC in RC cells and stimulated basal and 1alpha,25-(OH)(2)D(3)-stimulated PKC in GC cells, but wortmannin had no effect on the melittin-induced changes in either cell type. Pertussis toxin modestly increased the effect of 24R,25-(OH)(2)D(3) on PKC, whereas GDPbetaS had no effect, suggesting that PLD2 is the isoform responsible. This indicates that 1alpha,25-(OH)(2)D(3) regulates PKC in GC cells via PI-PLC and PLA(2), but not PC-PLC or PLD, whereas 24R,25-(OH)(2)D(3) regulates PKC in RC cells via PLD2.  相似文献   

8.
Boyan BD  Sylvia VL  Dean DD  Schwartz Z 《Steroids》2002,67(6):421-427
1 alpha,25(OH)(2)D(3) and 24R,25(OH)(2)D(3) mediate their effects on chondrocytes and osteoblasts in part through increased activity of protein kinase C (PKC). For both cell types, 1 alpha,25(OH)(2)D(3) exerts its effects primarily on more mature cells within the lineage, whereas 24R,25(OH)(2)D(3) exerts its effects primarily on relatively immature cells. Studies using the rat costochondral cartilage growth plate as a model indicate that the two metabolites increase PKC activity by different mechanisms. In growth zone cells (prehypertrophic/upper hypertrophic cell zones), 1 alpha,25(OH)(2)D(3) causes a rapid increase in PKC that does not involve new gene expression. 1 alpha,25(OH)(2)D(3) binds its membrane receptor (1,25-mVDR), resulting in activation of phospholipase A(2) and the rapid release of arachidonic acid, as well as activation of phosphatidylinositol-specific phospholipase C, resulting in formation of diacylglycerol and inositol-1,4,5-tris phosphate (IP(3)). IP(3) leads to release of intracellular Ca(2+) from the rough endoplasmic reticulum, and together with diacylglycerol, the increased Ca(2+) activates PKC. PKC is then translocated to the plasma membrane, where it initiates a phosphorylation cascade, ultimately phosphorylating the extracellular signal-regulated kinase-1 and -2 (ERK1/2) family of MAP kinases (MAPK). PKC increases are maximal at 9 min, and MAPK increases are maximal at 90 min in these cells. By contrast, 24R,25(OH)(2)D(3) increases PKC through activation of phospholipase D in resting zone cells. Peak production of diacylglycerol via phospholipase D2 is at 90 min, as are peak increases in PKC. Some of the effect is direct on existing plasma membrane PKC, but most is due to new PKC expression; translocation is not involved. Arachidonic acid and its metabolites also play differential roles in the mechanisms, stimulating PKC in growth zone cells and inhibiting PKC in resting zone cells. 24R,25(OH)(2)D(3) decreases phospholipase A(2) activity and prostaglandin production, thereby overcoming this potential inhibitory component, which may account for the delay in the PKC response. Ultimately, ERK1/2 is phosphorylated. PKC-dependent MAPK activity transduces some, but not all, of the physiological responses of each cell type to its respective vitamin D metabolite, suggesting that the membrane receptor(s) and nuclear receptor(s) may function interdependently to regulate proliferation and differentiation of musculoskeletal cells, but different pathways are involved at different stages of phenotypic maturation.  相似文献   

9.
10.
Matrix vesicles are extracellular organelles involved in mineral formation that are regulated by 1alpha,25(OH)(2)D(3). Prior studies have shown that protein kinase C (PKC) activity is involved in mediating the effects of 1alpha,25(OH)(2)D(3) in both matrix vesicles and plasma membranes. Here, we examined the regulation of matrix vesicle PKC by 1alpha,25(OH)(2)D(3) during biogenesis and after deposition in the matrix. When growth zone costochondral chondrocytes were treated for 9 min with 1alpha,25(OH)(2)D(3), PKCzeta in matrix vesicles was inhibited, while PKCalpha in plasma membranes was increased. In contrast, after treatment for 12 or 24 h, PKCzeta in matrix vesicles was increased, while PKCalpha in plasma membranes was unchanged. The effect of 1alpha,25(OH)(2)D(3) was stereospecific and metabolite-specific. Monensin blocked the increase in matrix vesicle PKC after 24 h, suggesting the secosteroid-regulated packaging of PKC. In addition, the 1alpha,25(OH)(2)D(3) membrane vitamin D receptor (1,25-mVDR) was involved, since a specific antibody blocked the 1alpha,25(OH)(2)D(3)-dependent changes in PKC after both long and short treatment times. In contrast, antibodies to annexin II had no effect, and there was no evidence for the presence of the nuclear VDR on Western blots. To investigate the signaling pathways involved in regulating matrix vesicle PKC activity after biosynthesis, matrix vesicles were isolated and then treated for 9 min with 1alpha,25(OH)(2)D(3) in the presence and absence of specific inhibitors. Inhibition of phosphatidylinositol-phospholipase C, phospholipase D, or G(i)/G(s) had no effect. However, inhibition of G(q) blocked the effect of 1alpha,25(OH)(2)D(3). The rapid effect of 1alpha,25(OH)(2)D(3) also involved the 1,25-mVDR. Moreover, arachidonic acid was found to stimulate PKC when added directly to isolated matrix vesicles. These results indicate that matrix vesicle PKC is regulated by 1alpha,25(OH)(2)D(3) at three levels: 1) during matrix vesicle biogenesis; 2) through direct action on the membrane; and 3) through production of other factors such as arachidonic acid.  相似文献   

11.
12.
13.
14.
The antiproliferative effect of 1alpha,25-dihydroxyvitamin D(3) (1alpha,25(OH)(2)D(3)) has been studied for a decade in diverse model systems, but the signalling pathways linking 1alpha,25(OH)(2)D(3) to cell cycle arrest remains unclear. In our attempt to establish a model system which would allow further identification of important players in the process of the 1alpha,25(OH)(2)D(3) imposed cell cycle arrest, we have isolated derivatives of the human breast cancer cell line MCF-7 and chosen two nearly 1alpha,25(OH)(2)D(3) resistant and two hypersensitive sub-clones. Investigation of cell cycle proteins regulated by 1alpha,25(OH)(2)D(3) in these clones indicates that activation of one component/pathway is responsible for the linkage between 1alpha,25(OH)(2)D(3) and growth arrest. Protein levels of the Vitamin D receptor (VDR) were elevated in sensitive cells upon 1alpha,25(OH)(2)D(3) treatment, whereas resistant clones were unable to induce VDR upon 1alpha,25(OH)(2)D(3) treatment. Our data show that VDR protein levels and the ability of a cell to induce VDR upon 1alpha,25(OH)(2)D(3) treatment correlate with the antiproliferative effects of 1alpha,25(OH)(2)D(3), and suggest that the level of VDR in cancer cells might serve as a prognostic marker for treatment of cancer with 1alpha,25(OH)(2)D(3) analogues.  相似文献   

15.
16.
17.
18.
Boyan BD  Sylvia VL  Dean DD  Schwartz Z 《Steroids》2001,66(3-5):363-374
The purpose of this paper is to summarize recent advances in our understanding of the physiological role of 24(R),25(OH)(2)D(3) in bone and cartilage and its mechanism of action. With the identification of a target cell, the growth plate resting zone (RC) chondrocyte, we have been able to use cell biology methodology to investigate specific functions of 24(R),25(OH)(2)D(3) and to determine how 24(R),25(OH)(2)D(3) elicits its effects. These studies indicate that there are specific membrane-associated signal transduction pathways that mediate both rapid, nongenomic and genomic responses of RC cells to 24(R),25(OH)(2)D(3). 24(R),25(OH)(2)D(3) binds RC chondrocyte membranes with high specificity, resulting in an increase in protein kinase C (PKC) activity. The effect is stereospecific; 24R,25(OH)(2)D(3), but not 24S,25-(OH)(2)D(3), causes the increase, indicating a receptor-mediated response. Phospholipase D-2 (PLD2) activity is increased, resulting in increased production of diacylglycerol (DAG), which in turn activates PKC. 24(R),25(OH)(2)D(3) does not cause translocation of PKC to the plasma membrane, but activates existing PKCalpha. There is a rapid decrease in Ca(2+) efflux, and influx is stimulated. 24(R),25(OH)(2)D(3) also reduces arachidonic acid release by decreasing phospholipase A(2) (PLA(2)) activity, thereby decreasing available substrate for prostaglandin production via the action of cyclooxygenase-1. PGE(2) that is produced acts on the EP1 and EP2 receptors expressed by RC cells to downregulate PKC via protein kinase A, but the reduction in PGE(2) decreases this negative feedback mechanism. Both pathways converge on MAP kinase, leading to new gene expression. One consequence of this is production of new matrix vesicles containing PKCalpha and PKCzeta and an increase in PKC activity. The chondrocytes also produce 24(R),25(OH)(2)D(3), and the secreted metabolite acts directly on the matrix vesicle membrane. Only PKCzeta is directly affected by 24(R),25(OH)(2)D(3) in the matrix vesicles, and activity of this isoform is inhibited. This effect may be involved in the control of matrix maturation and turnover. 24(R),25(OH)(2)D(3) causes RC cells to mature along the endochondral developmental pathway, where they become responsive to 1alpha,25(OH)(2)D(3) and lose responsiveness to 24(R),25(OH)(2)D(3), a characteristic of more mature growth zone (GC) chondrocytes. 1alpha,25(OH)(2)D(3) elicits its effects on GC through different signal transduction pathways than those used by 24(R),25(OH)(2)D(3). These studies indicate that 24(R),25(OH)(2)D(3) plays an important role in endochondral ossification by regulating less mature chondrocytes and promoting their maturation in the endochondral lineage.  相似文献   

19.
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