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 796 毫秒
1.
We have previously shown that 24,25-(OH)2D3 plays a major role in resting zone (RC) chondrocyte differentiation and that this vitamin D metabolite regulates protein kinase C (PKC). The aim of the present study was to identify the signal transduction pathway used by 24,25-(OH)2D3 to stimulate PKC activation. Confluent, fourth passage RC cells from rat costochondral cartilage were used to evaluate the mechanism of PKC activation. Treatment of RC cultures with 24,25-(OH)2D3 for 90 min produced a dose-dependent increase in diacylglycerol (DAG). Addition of R59022, a diacylglycerol kinase inhibitor, significantly increased PKC activity in cultures treated with 24,25-(OH)2D3. Addition of dioctanoylglycerol (DOG) to plasma membranes isolated from RC increased PKC activity 447-fold. Addition of pertussis toxin or cholera toxin to control cultures elevated basal PKC activity. When added together with 10−9 M 24,25-(OH)2D3, there was an additive effect on PKC activity but in cultures treated with 10−8 M 24,25-(OH)2D3, only the hormone-dependent stimulation of PKC was observed. The phospholipase C inhibitor, U73-122, had no effect on PKC activity, indicating that the DAG produced in response to 24,25-(OH)2D3 is not derived from phosphatidylinositol. Addition of the tyrosine kinase inhibitor, genistein, also had no effect on 24,25-(OH)2D3-stimulated PKC, further supporting the hypothesis that phospholipase C is not involved in the mechanism and that phospholipase D is responsible for the increase in DAG production. Phospholipase A2 inhibitors, quinacrine and AACOCF3, and the cyclooxygenase inhibitor indomethacin increased PKC activity in the RC cultures. Exogenous PGE2, one of the downstream products of phospholipase A2 action, inhibited PKC activity. These results suggest that 24,25-(OH)2D3 regulates PKC activity by two distinct phospholipid-dependent mechanisms: production of DAG via phospholipase D and inhibition of the production of PGE2 via inhibition of phospholipase A2 and cyclooxygenase. © 1996 Wiley-Liss, Inc.  相似文献   

2.
1,25-(OH)2D3 and 24,25-(OH)2D3 mediate their effects on chondrocytes through the classic vitamin D receptor (VDR) as well as through rapid membrane-mediated mechanisms which result in both nongenomic and genomic effects. In intact cells, it is difficult to distinguish between genomic responses via the VDR and genomic and nongenomic responses via membrane-mediated pathways. In this study, we used two hybrid analogues of 1,25-(OH)2D3 which have been modified on the A-ring and C,D-ring side chain (1α-(hydroxymethyl)-3β-hydroxy-20-epi-22-oxa-26,27-dihomo vitamin D3 (analogue MCW-YA = 3a) and 1β-(hydroxymethyl)-3α-hydroxy-20-epi-22-oxa-26,27-dihomo vitamin D3 (analogue MCW-YB = 3b) to examine the role of the VDR in response of rat costochondral resting zone (RC) and growth zone (GC) chondrocytes to 1,25-(OH)2D3 and 24,25-(OH)2D3. These hybrid analogues are only 0.1% as effective in binding to the VDR from calf thymus as 1,25-(OH)2D3. Chondrocyte proliferation ([3H]-thymidine incorporation), proteoglycan production ([35S]-sulfate incorporation), and activity of protein kinase C (PKC) were measured after treatment with 1,25-(OH)2D3, 24,25-(OH)2D3, or the analogues. Both analogues inhibited proliferation of both cell types, as did 1,25-(OH)2D3 and 24,25-(OH)2D3. Analogue 3a had no effect on proteoglycan production by GCs but increased that by RCs. Analogue 3b increased proteoglycan production in both GC and RC cultures. Both analogues stimulated PKC in GC cells; however, neither 3a nor 3b had an effect on PKC activity in RC cells. 1,25-(OH)2D3 and 3a decreased PKC in matrix vesicles from GC cultures, whereas plasma membrane PKC activity was increased, with 1,25-(OH)2D3 having a greater effect. 24,25-(OH)2D3 caused a significant decrease in PKC activity in matrix vesicles from RC cultures; 24,25-(OH)2D3, 3a, and 3b increased PKC activity in the plasma membrane fraction, however. Thus, with little or no binding to calf thymus VDR, 3a and 3b can affect cell proliferation, proteoglycan production, and PKC activity. The direct membrane effect is analogue-specific and cell maturation–dependent. By studying analogues with greatly reduced affinity for the VDR, we have provided further evidence for the existence of a membrane receptor(s) involved in mediating nongenomic effects of vitamin D metabolites. J. Cell. Biochem. 66:457–470, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

3.
1,25-(OH)2D3 (1,25) and 24,25-(OH)2D3 (24,25) mediate their effects on chondrocytes through the classic vitamin D receptor (VDR) as well as through rapid membrane-mediated mechanisms, which result in both nongenomic and genomic effects. In intact cells, it is difficult to distinguish between genomic responses via the VDR and genomic and nongenomic responses via membrane-mediated pathways. In this study, we used two analogues of 1,25 that have been modified on the A-ring (2a, 2b) and are only 0.1% as effective in binding to the VDR as 1,25, to examine the role of the VDR in the response of rat costochondral resting zone (RC) and growth zone (GC) chondrocytes to 1,25 and 24,25. Chondrocyte proliferation ([3H]-thymidine incorporation), proteoglycan production ([35S]-sulfate incorporation), and second messenger activation (activity of protein kinase C) were measured after treatment with 10-8 M 1,25, 10-7 M 24,25, or the analogues at 10-9–10-6 M. Both analogues inhibited proliferation of both cell types, as did 1,25 and 24,25. Neither 2a nor 2b had an effect on proteoglycan production by GCs or RCs. 2a caused a dose-dependent stimulation of protein kinase C (PKC) that was not inhibited by cycloheximide or actinomycin D in either GC or RC cells. 2b, on the other hand, had no effect on PKC activity in RCs and only a slight stimulatory effect in GCs. Both cells produce matrix vesicles, extracellular organelles associated with the initial stages of calcification, in culture that are regulated by vitamin D metabolites. Since these organelles contain no DNA or RNA, they provide an excellent model for studying the mechanisms used by vitamin D metabolites to mediate their nongenomic effects. When matrix vesicles were isolated from naive cultures of growth zone cells and treated with 2a, a dose-dependent inhibition of PKC activity was observed that was similar to that found with 1,25-(OH)2D3. Plasma membranes contained increased PKC activity after treatment with 2a, but the magnitude of the effect was less than that seen with 1,25-(OH)2D3. Analogue 2b had no affect on PKC activity in either membrane fraction. When matrix vesicles from resting zone chondrocyte cultures were treated with 24,25-(OH)2D3, a significant decrease in PKC activity was observed. No change in enzyme activity was found for either 1,25-(OH)2D3 or the analogues. PKC activity in the plasma membrane fraction, however, was increased by 24,25-(OH)2D3 as well as by analogue 2a. This study shows that these analogues, with little or no binding to the vitamin D receptor, can affect cell proliferation and PKC activity, but not proteoglycan production. The direct membrane effect is analogue specific and cell maturation dependent. Further, by eliminating the VDR-mediated component of the cellular response, we have provided further evidence for the existence of a membrane receptor(s) involved in mediating nongenomic effects of vitamin D metabolites. J. Cell. Physiol. 171:357–367, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

4.
Prior studies have shown that vitamin D regulation of protein kinase C activity (PKC) in the cell layer of chondrocyte cultures is cell maturation-dependent. In the present study, we examined the membrane distribution of PKC and whether 1α,25-(OH)2D3 and 24R,25-(OH)2D3 can directly regulate enzyme activity in isolated plasma membranes and extracellular matrix vesicles. Matrix vesicle PKC was activated by bryostatin-1 and inhibited by a PKC-specific pseudosubstrate inhibitor peptide. Depletion of membrane PKC activity using isoform-specific anti-PKC antibodies suggested that PKCα is the major isoform in cell layer lysates as well as in plasma membranes isolated from both cell types; PKCζ is the predominant form in matrix vesicles. This was confirmed in Western blots of immunoprecipitates as well as in studies using control peptides to block binding of the isoform specific antibody to the enzyme and using a PKCζ-specific pseudosubstrate inhibitor peptide. The presence of PKCζ in matrix vesicles was further verified by immunoelectron microscopy. Enzyme activity in the matrix vesicle was insensitive to exogenous lipid, whereas that in the plasma membrane required lipid for full activity. 1,25-(OH)2D3 and 24,25-(OH)2D3 inhibited matrix vesicle PKC, but stimulated plasma membrane PKC when added directly to the isolated membrane fractions. PKC activity in the matrix vesicle was calcium-independent, whereas that in the plasma membrane required calcium. Moreover, the vitamin D-sensitive PKC in matrix vesicles was not dependent on calcium, whereas the vitamin D-sensitive enzyme in plasma membranes was calcium-dependent. It is concluded that PKC isoforms are differentially distributed between matrix vesicles and plasma membranes and that enzyme activity is regulated in a membrane-specific manner. This suggests the existence of a nongenomic mechanism whereby the effects of 1,25-(OH)2D3 and 24,25-(OH)2D3 may be mediated via PKC. Further, PKCζ may be important in nongenomic, autocrine signal transduction at sites distal from the cell. © 1996 Wiley-Liss, Inc.  相似文献   

5.
Matrix metalloproteinases (MMPs) are a group of enzymes with the potential to degrade extracellular matrix proteins. One of the MMPs, stromelysin-1 (MMP-3) has been localized to extracellular matrix vesicles in growth plate chondrocyte cultures, suggesting involvement of this enzyme in remodeling of the extracellular matrix during endochondral development, a process which is regulated by the vitamin D metabolites, 1,25-(OH)2D3 and 24,25-(OH)2D3. To determine whether stromelysin-1 is regulated by vitamin D as well, confluent cultures of cells derived from growth zone (GC) and resting zone (RC) rat costochondral cartilage were treated with 1α,25-(OH)2D3 (1,25) and 24R,25-(OH)2D3 (24,25), respectively, and the effect on stromelysin-1 assessed by casein gel zymography and Western blots. Although stromelysin-1 activity was enriched in the matrix vesicle fraction, only the plasma membrane enzyme was affected by the treatment; 1,25 and 24,25 caused a marked decrease in plasma membrane stromelysin-1 activity in their target cells. Since plasma membrane protein kinase C (PKC) activity is stimulated by 1,25 and 24,25, we hypothesized that stromelysin-1 activity was regulated by the vitamin D metabolites via PKC-dependent phosphorylation. To test this, membrane fractions (containing endogenous PKCα and ζ as well as stromelysin-1) were incubated in the presence of purified rat brain PKC and/or recombinant human (rh) stromelysin-1 and [γ32P]-ATP and anti-stromelysin-1 immunoprecipitates were analyzed by autoradiography and Western blots. Immuno-phospho-stromelysin-1 was localized to a 52-kDa band in the plasma membrane fraction only; no phosphorylation was observed in the matrix vesicle fraction. Selective inhibitors of PKC activity demonstrated that phosphorylation was inhibited by H7 and low concentrations of H8, but not by HA1004, indicating that PKC, not PKA, was responsible. Protein phosphatase 2A, (PP2A), a serine/threonine-specific phosphatase, selectively removed the radiolabel in a time-dependent manner, providing further support for a PKC-dependent phosphorylation mechanism. Incubation of resting zone cell plasma membranes with 24,25, but not 1,25, resulted in phosphorylation of stromelysin-1, demonstrating that the nongenomic effect was metabolite-specific. This suggests that this may be one mechanism by which vitamin D metabolites regulate stromelysin-1 activity and that PKC-dependent phosphorylation inhibits the metalloproteinase. © 1996 Wiley-Liss, Inc.  相似文献   

6.
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   

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.
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   

9.
The effect of 24,25(OH)2D3 on 1,25(OH)2D3-induced hypercalcemia was studied in normal rats. Serum (S) levels and urinary excretion of Ca2+ (UCaV) were measured in (a) control rats, (b) rats receiving a daily sc injection of 54 ng 1,25(OH)2D3, (c) rats receiving 24,25(OH)2D3 in the same dose and same manner, and (d) rats receiving 1,25(OH)2D3 + 24,25(OH)2D3. The animals were housed in metabolic cages and 24-hr urine specimens were collected. After 24 hr SCa2+ increased similarly with 1,25(OH)2D3 and with 1,25(OH)2D3 + 24,25(OH)2D3, while 24,25(OH)2D3 alone did not change SCa2+. UCaV after 24 hr increased significantly less (P less than 0.025) with 1,25(OH)2D3 + 24,25(OH)2D3 than with 1,25(OH)2D3 alone. After 5 days of 1,25(OH)2D3, SCa2+ rose from 5.1 +/- 0.15 to 6.29 +/- 0.08 whereas 1,25(OH)2D3 + 24,25(OH)2D3 effected a greater increase in SCa2+ up to 6.63 +/- 0.09 (P less than 0.01). 24,25(OH)2D3 alone did not change SCa2+. UCaV after 5 days of treatment rose similarly with 1,25(OH)2D3 and with 1,25(OH)2D3 + 24,25(OH)2D3. After 10 days of 1,25(OH)2D3 SCa2+ was 6.17 +/- 0.15 meq/liter while with the combination SCa2+ rose to 6.74 +/- 0.2 (P less than 0.025). 24,25(OH)2D3 alone did not change SCa2+. These results show that (a) 24,25(OH)2D3 alone does not alter SCa2+ in normal rats, (b) combined administration of 1,25(OH)2D3 + 24,25(OH)2D3 enhances the hypercalcemic response to 1,25(OH)2D3 without a parallel increase in UCaV, and (c) it is suggested that the effect of 24,25(OH)2D3 on serum Ca2+ level, at least partly, may result from its hypocalciuric effect.  相似文献   

10.
11.
12.
The steroid derivative 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) is a regulator of bone biology, and there is evidence that 1,25(OH)2D3 modulates arachidonic acid metabolism in osteoblastic cell model systems and in bone organ cultures. In the present studies, 1,25(OH)2D3 decreased prostaglandin (PG) biosynthesis by normal adult human osteoblast-like (hOB) cell cultures by about 30%. The decrease was observed under basal incubation conditions, or in specimens stimulated by transforming growth factor-β1 (TGF-β) or by tumor necrosis factor-α (TNF). The inhibition of the TGF-β-stimulated PG production appeared to reflect a diminished efficiency of arachidonic acid conversion into PGs by the cells, while the efficiency of substrate utilization for PG biosynthesis was unaffected by 1,25(OH)2D3 pretreatment in the unstimulated samples, or in samples stimulated with TNF or with TNF plus TGF-β. Free arachidonic acid levels were decreased following 1,25(OH)2D3 pretreatment in the TNF stimulated samples. hOB cell phospholipase A2 activity was measured in subcellular fractions, and this activity was decreased by 20–25% in the 1,25(OH)2D3 pretreated samples. The addition of the selective inhibitor AACOCF3 to the phospholipase A2 assays provided evidence that it was the cytoplasmic isoform of the enzyme that was affected by the 1,25(OH)2D3 pretreatment of the hOB cells. Thus, 1,25(OH)2D3 regulation of hOB cell biology includes significant effects on arachidonic acid metabolism. In turn, this could influence the effects of other hormones and cytokines whose actions include the stimulated production of bioactive arachidonic acid metabolites. J. Cell. Biochem. 68:237–246, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

13.
Alkaline phosphatase activity appears to be altered when chondrocyte cultures are incubated with 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3). This study examined whether the hormone-responsive enzyme activity is associated with alkaline phosphatase-enriched extracellular membrane organelles called matrix vesicles. Confluent, third passage cultures of rat costochondral growth cartilage (GC) or resting zone chondrocytes (RC) were incubated with 1,25-(OH)2D3 or 24,25-dihydroxyvitamin D3 (24,25-(OH)2D3) and enzyme specific activity was assayed in the cell layer or in isolated matrix vesicle and plasma membrane fractions. Alkaline phosphatase-specific activity in the matrix vesicles was enriched at least 2-fold over that of the plasma membrane and 10-fold over that of the cell layer. Matrix vesicle alkaline phosphatase was stimulated by 1,25-(OH)2D3 in GC cultures and by 24,25-(OH)2D3 in RC cultures. The cell layer failed to reveal these subtle differences. 1,25-(OH)2D3 increased GC enzyme activity but the effect was one-half that observed in the matrix vesicles alone. No effect of 1,25-(OH)2D3 on enzyme activity of the RC cell layer or of 24,25-(OH)2D3 on either GC or RC cell layers was detected. Thus, response to the metabolites is dependent on chondrocytic differentiation and is site specific: the matrix vesicle fraction is targeted and not the cells per se.  相似文献   

14.
The effect of 24,25(OH)2D3 on 1,25(OH)2D3-induced hypercalcemia was studied in parathyroidectomized (PTX) rats for 10 days. Serum (S) and urinary Ca excretion (UCaV) were measured in (a) control rats, (b) rats receiving a daily sc injection of 54 ng 1,25(OH)2D3, (c) rats receiving 24,25(OH)2D3 in the same dose and same manner, and (d) rats receiving 1,25(OH)2D3 + 24,25(OH)2D3. Our results show that (i) 24,25(OH)2D3 alone does not increase SCa2+ in PTX rats, (ii) combined administration of 1,25(OH)2D3 + 24,25(OH)2D3 enhances the hypercalcemic response to 1,25(OH)2D3 without a parallel increase in UCaV, (iii) combined administration of 1,25(OH)2D3 + 24,25(OH)2D3 reduces the rise in urinary excretion of Ca2+ compared with that of rats receiving 1,25(OH)2D3 alone for 10 days, and (iv) these alterations are independent of parathyroid hormone.  相似文献   

15.
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   

16.
An adequate vitamin D status is essential to optimize muscle strength. However, whether vitamin D directly reduces muscle fiber atrophy or stimulates muscle fiber hypertrophy remains subject of debate. A mechanism that may affect the role of vitamin D in the regulation of muscle fiber size is the local conversion of 25(OH)D to 1,25(OH)2D by 1α‐hydroxylase. Therefore, we investigated in a murine C2C12 myoblast culture whether both 1,25(OH)2D3 and 25(OH)D3 affect myoblast proliferation, differentiation, and myotube size and whether these cells are able to metabolize 25(OH)D3 and 1,25(OH)2D3. We show that myoblasts not only responded to 1,25(OH)2D3, but also to the precursor 25(OH)D3 by increasing their VDR mRNA expression and reducing their proliferation. In differentiating myoblasts and myotubes 1,25(OH)2D3 as well as 25(OH)D3 stimulated VDR mRNA expression and in myotubes 1,25(OH)2D3 also stimulated MHC mRNA expression. However, this occurred without notable effects on myotube size. Moreover, no effects on the Akt/mTOR signaling pathway as well as MyoD and myogenin mRNA levels were observed. Interestingly, both myoblasts and myotubes expressed CYP27B1 and CYP24 mRNA which are required for vitamin D3 metabolism. Although 1α‐hydroxylase activity could not be shown in myotubes, after treatment with 1,25(OH)2D3 or 25(OH)D3 myotubes showed strongly elevated CYP24 mRNA levels compared to untreated cells. Moreover, myotubes were able to convert 25(OH)D3 to 24R,25(OH)2D3 which may play a role in myoblast proliferation and differentiation. These data suggest that skeletal muscle is not only a direct target for vitamin D3 metabolites, but is also able to metabolize 25(OH)D3 and 1,25(OH)2D3. J. Cell. Physiol. 231: 2517–2528, 2016. © 2016 The Authors. Journal of Cellular Physiology Published by Wiley Periodicals, Inc.  相似文献   

17.
Summary After injection of radiolabeled 1,25 (OH)2 vitamin D3, nuclear concentration of radioactivity is observed in parenchymal cells of the parathyroid gland in pregnant, adult male, and 10-day male neonatal rats. In competition studies with unlabeled 1,25 (OH)2 vitamin D3, but not with 25 (OH) vitamin D3, nuclear uptake is prevented. Experiments with 3H 25 (OH) vitamin D3, in contrast to 3H 1,25 (OH)2 vitamin D3, do not show nuclear concentration in cells of the parathyroid. The results of the autoradiographic studies suggest the presence of receptors for a direct effect of 1,25 (OH)2 vitamin D3 on the parathyroid gland for modulation of parathyroid hormone secretion.  相似文献   

18.
The present study was undertaken to evaluate the effect of 24,25(OH)2D3 on serum calcium concentration in rats with reduced renal mass. Adult 5/6 nephrectomized male rats were divided into four groups: (i) control rats, (ii) rats treated with 1,25(OH)2D3, (iii) rats treated with 24,25(OH)2D3, and (iv) rats treated with 1,25(OH)2D3 and 24,25(OH)2D3. After 4 days, serum calcium in the 1,25(OH)2D3-treated group was 7.13 +/- 0.32 meq/liter (P less than 0.001 vs control). With the combination of 1,25(OH)2D3 and 24,25(OH)2D3 serum calcium was higher than that in control, 6.25 +/- 0.5 meq/liter (P less than 0.001 vs control), but lower than that in rats receiving 1,25(OH)2D3 alone (P less than 0.05). No change in serum calcium was seen in animals treated with 24,25(OH)2D3 alone. On the eighth day serum calcium in the 1,25(OH)2D3-treated group, 6.52 +/- 0.25, was higher than in the 1,25(OH)2D3 + 24,25(OH)2D3 group, 5.87 +/- 0.17 meq/liter, P less than 0.05, P less than 0.001 vs control. In both 1,25(OH)2D3- and 1,25(OH)2D3 + 24,25(OH)2D3-treated rats, hypercalciuria of similar magnitude occurred on the fourth and eighth day of treatment. No change in urinary calcium was seen in the control and 24,25(OH)2D3-treated rats. Thus, in 5/6 nephrectomized rats combined administration of 1,25(OH)2D3 and 24,25(OH)2D3 attenuates the calcemic response to 1,25(OH)2D3 without changes in urinary calcium excretion. These observations suggest that the effect of 24,25(OH)2D3 on serum calcium is different in 5/6 nephrectomized rats as compared to normal rats, in which an augmentation of serum calcium was observed following administration of both vitamin D metabolites. The effect of 24,25(OH)2D3 on serum calcium in rats with reduced renal mass may result from a direct effect of 24,25(OH)2D3 on the bone.  相似文献   

19.
1,25-(OH)2D3 (1,25) exerts its effects on growth plate chondrocytes through classical vitamin D (VDR) receptor-dependent mechanisms, resulting in mineralization of the extracellular matrix. Recent studies have shown that membrane-mediated mechanisms are involved as well. 1,25 targets cells in the prehypertrophic and upper hypertrophic zones of the costochondral cartilage growth plate (GC cells), resulting in increased specific activity of alkaline phosphatase (ALP), phospholipase A2 (PLA2), and matrix metalloproteinases (MMPs). At the cellular level, 1,25 action results in rapid changes in arachidonic acid (AA) release and re-incorporation, alterations in membrane fluidity and Ca ion flux, and increased prostaglandin E1 and E2 (PGE2) production. Protein kinase C (PKC) is activated in a phospholipase C (PLC) dependent-mechanism, due in part to the increased production of diacylglycerol (DAG). In addition, AA acts directly on the cell to increase PKC specific activity. AA also provides a substrate for cyclooxygenase (COX), resulting in PGE2 production. 1,25 mediates its effects through COX-1, the constitutive enzyme, but not COX-2, the inducible enzyme. Time course studies using specific inhibitors of COX-1 show that AA stimulates PKC activity and PKC then stimulates PGE2 production. PGE2 acts as a mediator of 1,25 action on the cells, also stimulating PKC activity. The rapid effects of 1,25 on PKC are nongenomic, occurring within 3 min and reaching maximal activation by 9 min. It promotes translocation of PKC to the plasma membrane. When 1,25 is incubated directly with isolated plasma membranes, PKCalpha is stimulated although PKCzeta is also present. In contrast, when isolated matrix vesicles (MVs) are incubated with 1,25, PKCzeta is inhibited and PKCalpha is unaffected. These membrane-mediated effects are due to the presence of a specific membrane vitamin D receptor (mVDR) that is distinct from the classical cytosolic VDR. Studies using 1,25 analogs with reduced binding affinity for the classical VDR, confirm that rapid activation of PKC by 1,25 is not VDR dependent. The membrane-mediated effects of 1,25 are critical to the regulation of events in the extracellular matrix produced by the chondrocytes. MVs are extracellular organelles associated with maturation of the matrix, preparing it for mineralization. MV composition is under genomic control, involving VDR-mechanisms. In the matrix, no new gene expression or protein synthesis can occur, however. Differential distribution of PKC isoforms and their nongenomic regulation by 1,25 is one way for the chondrocyte to control events at sites distant from the cell. GC cells contain 1a-hydroxylase and produce 1,25; this production is regulated by 1,25, 24,25, and dexamethasone. 1,25 stimulates MMPs in the MVs, resulting in increased proteoglycan degradation in mineralization gels, and increased activation of latent transforming growth factor-beta 1 (TGF-beta1).  相似文献   

20.
The T47D human breast cancer cell line contains a specific binding protein for 1.25-(OH)2D3, with 15000 sites per cell. The Kd (1.1 × 10?10 M) and sedimentation coefficient on sucrose gradients (3.7S) are the same as those reported for the 1,25-(OH)2D3 receptor in other tissues. Other vitamin D3 metabolites bound to the receptor with an order of affinities 1,25-(OH)2D3 > 1,24,25-(OH)3D3 > 25-OHD3 > 24,25-(OH)2D3 > D3. A new analogue 1β,25-(OH)2D3 was only as effective as 24,25-(OH)2D3 at displacing the hormone from the receptor. Cell growth was stimulated in a dose dependent manner by the addition of 1,25-(OH)2D3 (up to 0.8 nM) to the medium. A higher concentration of hormone was without effect.  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