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.
Boyan BD  Wang L  Wong KL  Jo H  Schwartz Z 《Steroids》2006,71(4):286-290
1,25-Dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)] acts on chondrocytes and osteoblasts through traditional nuclear Vitamin D receptor (VDR) mechanisms as well as through rapid actions on plasma membranes that initiate intracellular signaling pathways. We have investigated the mechanisms involved in activation of protein kinase C (PKC) and downstream biological responses that depend on the latter pathway. These studies show that PKC activation depends on presence of a membrane receptor ERp60 and rapid increases in phospholipase A(2) (PLA(2)) activity. Cells that are responsive to 1alpha,25(OH)(2)D(3) express PLA(2) activating protein (PLAA), suggesting a link between ERp60 and PLA(2). Increased PLA(2) results in increased arachidonic acid release and formation of lysophospholipid, which then activates phospholipase C beta (PLCbeta), leading to rapid formation of inositol-trisphosphate (IP3) and diacylglycerol (DAG). PLA(2), PLC, and DAG are all associated with lipid rafts including caveolae in many cells, suggesting that the caveolar environment may be an important mediator of PKC activation by 1alpha,25(OH)(2)D(3). Here, we use the VDR(-/-) mouse costochondral cartilage growth plate to examine the expression of ERp60 and PLAA in vivo in 1alpha,25(OH)(2)D(3)-responsive hypertrophic chondrocytes (growth zone cells) and in resting zone cells that do not respond to this Vitamin D metabolite in vitro. In addition, we determined if intact lipid rafts are required for the response of rat costochondral cartilage growth zone cells to 1alpha,25(OH)(2)D(3). The results show that ERp60 and PLAA are localized to 1alpha,25(OH)(2)D(3)-responsive growth zone cells and metaphyseal osteoblasts, even in VDR(-/-) mice. Disruption of lipid rafts using beta-cyclodextrin blocks the activation of PKC by 1alpha,25(OH)(2)D(3) and reduces the ability of 1alpha,25(OH)(2)D(3) to regulate [(35)S]-sulfate incorporation.  相似文献   

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 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   

4.
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   

5.
Recent studies have shown that 24R,25-(OH)(2)D(3) mediates its effects on growth plate chondrocytes via membrane receptors. This study examined the roles of phospholipase A(2) (PLA(2)) and cyclooxygenase (Cox) in the mechanism of action of 24R, 25-(OH)(2)D(3) in resting zone chondrocytes in order to determine whether the activity of one or both enzymes provides a regulatory checkpoint in the signaling pathway resulting in increased protein kinase C (PKC) activity. We also determined whether constitutive or inducible Cox is involved. Cultures were incubated with 24R, 25-(OH)(2)D(3) for 90 min to measure PKC or for 24 h to measure physiological responses ([(3)H]-thymidine incorporation, alkaline phosphatase-specific activity, [(35)S]-sulfate incorporation). Based on RT-PCR and Northern blot analysis, resting zone chondrocytes express mRNAs for both Cox-1 and Cox-2. Levels of mRNA for both proteins were unchanged from control levels after a 24-h incubation with 24R,25-(OH)(2)D(3). To examine the role of Cox, the cultures were also treated with resveratrol (a specific inhibitor of Cox-1), NS-398 (a specific inhibitor of Cox-2), or indomethacin (a general Cox inhibitor). Cox-1 inhibition resulted in effects on proliferation, differentiation, and matrix production typical of 24R, 25-(OH)(2)D(3). In contrast, inhibition of Cox-2 had no effect, indicating that 24R,25-(OH)(2)D(3) exerts its effects via Cox-1. Inhibition of Cox-1 also blocked 24R,25-(OH)(2)D(3)-dependent increases in PKC. Activation of PLA(2) with melittin inhibited 24R, 25-(OH)(2)D(3)-dependent stimulation of PKC, and inhibition of PLA(2) with quinacrine stimulated PKC in response to 24R, 25-(OH)(2)D(3). Inclusion of resveratrol reduced the melittin-dependent inhibition of PLA(2) and caused an increase in quinacrine-stimulated PLA(2) activity. Metabolism of arachidonic acid to leukotrienes is not involved in the response to 24R, 25-(OH)(2)D(3) because inhibition of lipoxygenase had no effect. The effect of 24R,25-(OH)(2)D(3) was specific because 24S,25-(OH)(2)D(3), the biologically inactive stereoisomer, failed to elicit a response from the cells. These results support the hypothesis that 24R, 25-(OH)(2)D(3) exerts its effects via more than one signaling pathway and that these pathways are interrelated via the modulation of PLA(2). PKC regulation may occur at multiple stages in the signal transduction cascade.  相似文献   

6.
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   

7.
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   

8.
9.
Xie Z  Bikle DD 《Steroids》2001,66(3-5):339-345
Phospholipase C-gamma1 (PLC-gamma1) is the most abundant member of the phospholipase C family expressed in human keratinocytes. PLC-gamma1 is induced by 1alpha,25-dihydroxyvitamin D(3) (1alpha,25(OH)(2)D(3)) in normal keratinocytes via a DR6-type vitamin D responsive element. This regulation is not observed in transformed keratinocytes. The role of PLC-gamma1 in mediating 1alpha,25(OH)(2)D(3) and calcium-regulated differentiation was then tested. Both specific PLC inhibitors and antisense constructs which selectively block PLC-gamma1 production prevented 1alpha,25(OH)(2)D(3) and calcium from inducing markers of differentiation such as involucrin and transglutaminase. These studies demonstrate that PLC-gamma1 induction by 1alpha,25(OH)(2)D(3) is critical to the ability of this hormone to regulate keratinocyte differentiation.  相似文献   

10.
Transforming growth factor beta-1 (TGF-beta1) is released from the extracellular matrix of rat growth plate chondrocytes and activated by stromelysin-1 (matrix metalloproteinase 3, MMP-3), an enzyme that is stored in matrix vesicles. MMP-3 is released from these extracellular organelles by the direct action of 1alpha,25(OH)2D3 via activation of phospholipase A2 (PLA2), resulting in local production of lysophospholipids and matrix vesicle membrane destabilization. This effect of 1alpha,25(OH)2D3 is greater in matrix vesicles from growth zone chondrocyte cultures and PLA2 activity is higher in the growth zone in vivo, suggesting that it may depend on chondrocyte maturation state in the endochondral lineage. Previous studies have shown that latent TGF-beta1 can be activated by mild detergents in vitro, suggesting that lysophospholipids may act in vivo in a similar manner. To test this hypothesis, we determined if rat costochondral growth plate cartilage cells produce lysophosphatidylcholine (LPC) and lysophosphatidylethanolamine (LPE) in a maturation state-dependent manner and if LPC or LPE could release and activate latent TGF-beta1 from the extracellular matrix produced by these cells. Rat growth plate chondrocytes produced both lysophospholipids, with growth zone cells producing higher levels of LPE via PLA1, and resting zone cells producing higher levels of LPC via PLA2. LPC and LPE directly increased activation of recombinant human latent TGF-beta1 in a biphasic manner with a peak at 2 microg/ml. Phosphatidylcholine, phosphatidylethanolamine, and LPE plasmalogen (LPEP), but not choline, also activated TGF-beta1. Latent TGF-beta1 incubated with LPC or LPE, but neither lysophospholipid alone, stimulated [3H]-thymidine incorporation of resting zone cells, indicating the TGF-beta1 released was biologically active. LPC and LPE also released TGF-beta1 in a dose- and time-dependent manner when incubated with cell-free extracellular matrices produced by the cells. These results indicate that LPC and LPE have important roles as regulators of rat growth plate chondrocytes by directly and indirectly activating TGF-beta1 stored in the extracellular matrix.  相似文献   

11.
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   

12.
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   

13.
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   

14.
Many of the effects of 1α,25-(OH)2D3 and 24R,25-(OH)2D3 on costochondral chondrocytes are mediated by the protein kinase C (PKC) signal transduction pathway. 1α,25-(OH)2D3 activates PKC in costochondral growth zone chondrocytes through a specific membrane receptor (1α,25-mVDR), involving rapid increases in diacylglycerol via a phospholipase C (PLC)-dependent mechanism. 24R,25-(OH)2D3 activates PKC in resting zone chondrocytes. Although diacylglycerol is increased by 24R,25-(OH)2D3, PLC is not involved, suggesting a phospholipase D (PLD)-dependent mechanism. Here, we show that resting zone and growth zone cells express mRNAs for PLD1a, PLD1b, and PLD2. Both cell types have PLD activity, but levels are higher in resting zone cells. 24R,25-(OH)2D3, but not 24S,25-(OH)2D3 or 1α,25-(OH)2D3, stimulates PLD activity in resting zone cells within 3 min via nongenomic mechanisms. Neither 1α,25-(OH)2D3 nor 24R,25-(OH)2D3 affected PLD in growth zone cells. Basal and 24R,25-(OH)2D3-stimulated PLD were inhibited by the PLD inhibitors wortmannin and EDS. Inhibition of phosphatidylinositol 3-kinase (PI 3-kinase), PKC, phosphatidylinositol-specific PLC (PI-PLC), and phosphatidylcholine-specific PLC (PC-PLC) had no effect on PLD activity. Thus, 24R,25-(OH)2D3 stimulates PLD, and PI 3-kinase, PI-PLC and PKC are not involved, whereas PLD is required for stimulation of PKC by 24R,25-(OH)2D3. Pertussis toxin, GDPβS, and GTPγS had no effect on 24R,25-(OH)2D3-dependent PLD when added to cell cultures, indicating that G-proteins are not involved. These data show that PKC activation in resting zone cells is mediated by PLD and suggest that a functional 24R,25-(OH)2D3-mVDR is required. The results also support the conclusion that the 24R,25-(OH)2D3-responsive PLD is PLD2, since this PLD isoform is G-protein-independent.  相似文献   

15.
Recently, 25-hydroxyvitamin D3-24-hydroxylase (CYP24A1) has been shown to catalyze not only hydroxylation at C-24 but also hydroxylations at C-23 and C-26 of the secosteroid hormone 1alpha, 25-dihydroxyvitamin D3 (1alpha,25(OH)2D3). It remains to be determined whether CYP24A1 has the ability to hydroxylate vitamin D3 compounds at C-25. 1alpha,24(R)-dihydroxyvitamin D3 (1alpha,24(R)(OH)2D3) is a non-25-hydroxylated synthetic vitamin D3 analog that is presently being used as an antipsoriatic drug. In the present study, we investigated the metabolism of 1alpha,24(R)(OH)2D3 in human keratinocytes in order to examine the ability of CYP24A1 to hydroxylate 1alpha,24(R)(OH)2D3 at C-25. The results indicated that keratinocytes metabolize 1alpha,24(R)(OH)2D3 into several previously known both 25-hydroxylated and non-25-hydroxylated metabolites along with two new metabolites, namely 1alpha,23,24(OH)3D3 and 1alpha,24(OH)2-23-oxo-D3. Production of the metabolites including the 25-hydroxylated ones was detectable only when CYP24A1 activity was induced in keratinocytes 1alpha,25(OH)2D3. This finding provided indirect evidence to indicate that CYP24A1 catalyzes C-25 hydroxylation of 1alpha,24(R)(OH)2D3. The final proof for this finding was obtained through our metabolism studies using highly purified recombinant rat CYP24A1 in a reconstituted system. Incubation of this system with 1alpha,24(R)(OH)2D3 resulted in the production of both 25-hydroxylated and non-25-hydroxylated metabolites. Thus, in our present study, we identified CYP24A1 as the main enzyme responsible for the metabolism of 1alpha,24(R)(OH)2D3 in human keratinocytes, and provided unequivocal evidence to indicate that the multicatalytic enzyme CYP24A1 has the ability to hydroxylate 1alpha,24(R)(OH)2D3 at C-25.  相似文献   

16.
17.
1 alpha,25-(OH)(2)D(3) exerts its effects on chondrocytes and enterocytes via nuclear receptors (1,25-nVDR) and a separate membrane receptor (1,25-mVDR) that activates protein kinase C (PKC). 24R,25-(OH)(2)D(3) also stimulates PKC in chondrocytes, but through other membrane mechanisms. This study examined the hypothesis that osteoblasts possess distinct membrane receptors for 1 alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) that are involved in the activation of PKC and that receptor expression varies as a function of cell maturation state. 1 alpha,25-(OH)(2)D(3) stimulated PKC in well differentiated (UMR-106, MC-3T3-E1) and moderately differentiated (ROS 17/2.8) osteoblast-like cells, and in cultures of fetal rat calvarial (FRC) cells and 2T3 cells treated with rhBMP-2 to promote differentiation. 24R,25-(OH)(2)D(3) stimulated PKC in FRC and 2T3 cultures that had not been treated to induce differentiation, and in ROS 17/2.8 cells. MG63 cells, a relatively undifferentiated osteoblast-like cell line, had no response to either metabolite. Ab99, a polyclonal antibody generated to the chick enterocyte 1,25-mVDR, but not a specific antibody to the 1,25-nVDR, inhibited response to 1 alpha,25-(OH)(2)D(3). 1 alpha,25-(OH)(2)D(3) exhibited specific binding to plasma membrane preparations from cells demonstrating a PKC response to this metabolite that is typical of positive cooperativity. Western blots of these membrane proteins reacted with Ab99, and the Ab99-positive protein had an Mr of 64 kDa. There was no cross-reaction with antibodies to the C- or N-terminus of annexin II. The effect of 24,25-(OH)(2)D(3) on PKC was stereospecific; 24S,25-(OH)(2)D(3) had no effect. These results demonstrate that response to 1 alpha,25-(OH)(2)D(3) and 24R,25-(OH)(2)D(3) depends on osteoblast maturation state and suggest that specific and distinct membrane receptors are involved.  相似文献   

18.
The antiproliferative effect of 1alpha,25(OH)(2)D(3) on human prostate cancer cells is well known, but the mechanism is still not fully understood, especially its androgen-dependent action. Based on cDNA microarray results, we found that long-chain fatty-acid-CoA ligase 3 (FACL3/ACS3) might play an important role in vitamin D(3) and androgen regulation of LNCaP cell growth. The expression of FACL3/ACS3 was found to be significantly upregulated by 1alpha,25(OH)(2)D(3) and the regulation was shown to be time-dependent, with the maximal regulation over 3.5-fold at 96h. FACL3/ACS3 was a dominant isoform of FACL/ACS expressed in LNCaP cells as indicated by measuring the relative expression of each isoform. 1alpha,25(OH)(2)D(3) had no significant effect on the expression of FACL1(FACL2), FACL4 and FACL6 except for its downregulation of FACL5 at 24 and 48h by around twofold. The upregulation of FACL3/ACS3 expression by 1alpha,25(OH)(2)D(3) was accompanied with increased activity of FACL/ACS as demonstrated by enzyme activity assay using a (14)C-labeled substrate preferential for FACL3/ACS3. The growth inhibitory effect of 1alpha,25(OH)(2)D(3) on LNCaP cells was significantly attenuated by FACL3/ACS3 activity inhibitor. Androgen withdrawal (DCC-serum), in the presence of antiandrogen Casodex or in AR-negative prostate cancer cells (PC3 and DU145), vitamin D(3) failed to regulate FACL3/ACS3 expression. The upregulation of FACL3/ACS3 expression by vitamin D(3) was recovered by the addition of DHT in DCC-serum medium. Western blot analysis showed that the expression of androgen receptor (AR) protein was consistent with vitamin D(3) regulation of FACL3/ACS3 expression. Taken together, the data suggest that the upregulation of FACL3/ACS3 expression by vitamin D(3) is through an androgen/AR-mediated pathway and might be one of the contributions of the vitamin D(3) antiproliferative effect in prostate cancer LNCaP cells.  相似文献   

19.
During endochondral development, growth plate chondrocytes must remodel their matrix in a number of ways as they differentiate and mature. In previous studies, we have shown that matrix metalloproteinases (MMPs) extracted from matrix vesicles can extensively degrade aggrecan and that this is modulated by vitamin D metabolites in a manner involving protein kinase C (PKC). Matrix vesicles represent only a small component of the extracellular matrix, however, and it is unknown if the total metalloproteinase complement, including the MMPs and aggrecanases in the culture, is also regulated in a similar way. This study tested the hypothesis that vitamin D metabolites regulate the level of metalloproteinase activity in growth plate chondrocytes via a PKC-dependent mechanism and play a role in partitioning this proteinase activity between the media and cell layer (cells+matrix) in these cultures. To do this, resting zone cells (RC) were treated with 10(-9)-10(-7) M 24R,25-(OH)(2)D(3), while growth zone cells (GC) were treated with 10(-10)-10(-8) M 1alpha,25-(OH)(2)D(3). Cultures of both cell types were also treated with the PKC inhibitor chelerythrine in the presence and absence of vitamin D metabolites. At harvest, the media were either left untreated or treated to destroy metalloproteinase inhibitors, while enzyme activity in the cell layers was extracted with buffered guanidine and then treated like the media to destroy metalloproteinase inhibitors. Neutral metalloproteinase (aggrecan-degrading activity) activity was assayed on aggrecan-containing polyacrylamide gel beads and collagenase activity was measured on telopeptide-free type I collagen. Neutral metalloproteinase activity was found primarily in the cell layer of both cell types; however, activity was greater in extracts of GC cell layers. No collagenase activity could be detected in RC extracts until the metalloproteinase inhibitors were destroyed. In contrast, extracts of GC cell layers contained measurable activity without removing the inhibitors, and destroying the inhibitors resulted in a greater than two-fold increase in activity. No collagenase activity was found in the media of either cell type. 24,25-(OH)(2)D(3) caused a dose-dependent increase in neutral metalloproteinase activity in extracts of RC cells, but had no effect on collagenase activity. In contrast, 1,25-(OH)(2)D(3) caused a dose-dependent decrease in collagenase activity in extracts of GC cells, but had no effect on neutral metalloproteinase activity. In both cases, the effect of the vitamin D metabolite was mediated through the activation of PKC. These results support the hypothesis that metalloproteinases are involved in regulating the bulk turnover of collagen and aggrecan in growth plate chondrocytes and that the amount of metalloproteinase activity found is a function of the cell maturation state. Furthermore, 83-93% of neutral metalloproteinase activity and 100% of collagenase activity is localized to the cell layer. Moreover, the regulation of metalloproteinase activity by 1,25-(OH)(2)D(3) and 24,25-(OH)(2)D(3) involves a PKC-dependent pathway that is controlled by the target cell-specific vitamin D metabolite.  相似文献   

20.
We have previously shown that stimulation of proliferation of avian embryonic muscle cells (myoblasts) by 1alpha,25(OH)(2)-vitamin D(3) (1alpha,25(OH)(2)D(3)) is mediated by activation of the mitogen-activated protein kinase (MAPK; ERK1/2). To understand how 1alpha,25(OH)(2)D(3) up-regulates the MAPK cascade, we have investigated whether the hormone acts upstream through stimulation of Raf-1 and the signaling mechanism by which this effect might take place. Treatment of chick myoblasts with 1alpha,25(OH)(2)D(3) (1 nm) caused a fast increase of Raf-1 serine phosphorylation (1- and 3-fold over basal at 1 and 2 min, respectively), indicating activation of Raf-1 by the hormone. These effects were abolished by preincubation of cells with a specific Ras inhibitor peptide that involves Ras in 1alpha,25(OH)(2)D(3) stimulation of Raf-1. 1alpha,25(OH)(2)D(3) rapidly induced tyrosine de-phosphorylation of Ras-GTPase-activating protein, suggesting that inhibition of Ras-GTP hydrolysis is part of the mechanism by which 1alpha,25(OH)(2)D(3) activates Ras in myoblasts. The protein kinase C (PKC) inhibitors calphostin C, bisindolylmaleimide I, and Ro 318220 blocked 1alpha,25(OH)(2)D(3)-induced Raf-1 serine phosphorylation, revealing that hormone stimulation of Raf-1 also involves PKC. In addition, transfection of muscle cells with an antisense oligodeoxynucleotide against PKCalpha mRNA suppressed serine phosphorylation by 1alpha,25(OH)(2)D(3). The increase in MAPK activity and tyrosine phosphorylation caused by 1alpha,25(OH)(2)D(3) could be abolished by Ras inhibitor peptide, compound PD 98059, which prevents the activation of MEK by Raf-1, or incubation of cell lysates before 1alpha,25(OH)(2)D(3) exposure with an anti-Raf-1 antibody. In conclusion, these results demonstrate for the first time in a 1alpha,25(OH)(2)D(3) target cell that activation of Raf-1 via Ras and PKCalpha-dependent serine phosphorylation plays a central role in hormone stimulation of the MAPK-signaling pathway leading to muscle cell proliferation.  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