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 93 毫秒
1.
2.
The effects of 1alpha,25-dihydroxyvitamin D(3) [1,25(OH)(2)D(3)] are mainly mediated by nuclear receptors modulating gene expression. However, there are increasing evidences of nongenomic mechanisms of this hormone associated with kinase- and calcium-activated signaling pathways. In this context, the aim of the present work was to investigate the signaling pathways involved in the mechanism of action of 1,25(OH)(2)D(3) on vimentin phosphorylation in 15-day-old rat testes. Results showed that 1,25(OH)(2)D(3) at concentrations ranging from 1 nM to 1 microM increased vimentin phosphorylation independent of protein synthesis. We also demonstrated that the mechanisms underlying the hormone action involve protein kinase C activation in a phospholipase C-independent manner. Moreover, we showed that the participation of protein kinase A, extracellular regulated protein kinase (ERK), and intra- and extracellular Ca(2+) mediating the effects of 1,25(OH)(2)D(3) on the cytoskeleton. In addition, we investigated the effect of different times of exposure to the hormone on total and phosphoERK1/2 or c-Jun N-terminal kinases 1/2 (JNK1/2) in immature rat testis. Results showed that the total levels of ERK1/2 and JNK1/2 were unaltered from 1 to 15 min exposure to 1,25(OH)(2)D(3). However, the phosphoERK1/2 levels significantly increased at 1 and 5 min 1,25(OH)(2)D(3) treatment. Furthermore, phosphoJNK1 levels were decreased at 10 and 15 min 1,25(OH)(2)D(3) exposure, while phosphoJNK 2 levels were diminished at 5, 10 and 15 min treatment with the hormone. These findings demonstrate that 1,25(OH)(2)D(3) may modulate vimentin phosphorylation through nongenomic Ca(2+)-dependent mechanisms in testis cells.  相似文献   

3.
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   

4.
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   

5.
The steroid hormone 1,25-dihydroxyvitamin D3 [1,25(OH)2D3] rapidly stimulates the uptake of phosphate in isolated chick intestinal cells, while the steroid 24,25-dihydroxyvitamin D3 [24,25(OH)2D3] inhibits the rapid stimulation by 1,25(OH)2D3. Earlier work in this laboratory has indicated that a cellular binding protein for 24,25(OH)2D3 is the enzyme catalase. Since binding resulted in decreased catalase activity and increased H2O2 production, studies were undertaken to determine if pro-oxidant conditions mimicked the inhibitory actions of 24,25(OH)2D3, and anti-oxidant conditions prevented the inhibitory actions of 24,25(OH)2D3. An antibody against the 24,25(OH)2D3 binding protein was found to neutralize the inhibitory effect of the steroid on 1,25(OH)2D3-mediated 32P uptake. Incubation of cells in the presence of 50 nM catalase was also found to alleviate inhibition. In another series of experiments, isolated intestinal epithelial cells were incubated as controls or with 1,25(OH)2D3, each in the presence of the catalase inhibitor 3-amino-1,2,4-triazole, or with 1,25(OH)2D3 alone. Cells exposed to hormone alone again showed an increased accumulation of 32P, while cells treated with catalase inhibitor and hormone had uptake levels that were indistinguishable from controls. We tested whether inactivation of protein kinase C (PKC), the signaling pathway for 32P uptake, occurred. Incubation of cells with phorbol-13-myristate (PMA) increased 32P uptake, while cells pretreated with 50 microM H2O2 prior to PMA did not exhibit increased uptake. Likewise, PMA significantly increased PKC activity while cells exposed to H2O2 prior to PMA did not. It is concluded that catalase has a central role in mediating rapid responses to steroid hormones.  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.
10.
Farach-Carson MC 《Steroids》2001,66(3-5):357-361
1alpha,25-Dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)] treatment of osteoblastic cells elicits a series of measurable responses that include both rapid, membrane-initiated effects and longer-term nuclear receptor-mediated effects. Structural analogs have been identified and characterized that selectively activate subsets of these pathways. Two analogs from over 35 that have been tested were chosen for this comparison because they activate non-overlapping response pathways, presumably representing either membrane-initiated or nuclear receptor-initiated activities. Compound AT [25(OH)-16ene-23yne-D(3)] lacks the 1-hydroxyl essential for interacting with the nuclear receptor, but triggers Ca(2+) influx through plasma membrane Ca(2+) channels, augments parathyroid hormone (PTH)-induced Ca(2+) signals, dephosphorylates the matrix protein osteopontin (OPN), and along with PTH stimulates release of calcium from calvaria in organ culture. Compound BT [1alpha,24(OH)(2)-22ene-24cyclopropyl-D(3)] does not elicit any of the rapid responses or enhance PTH-induced bone resorption, but binds to the nuclear receptor for 1alpha,25(OH)(2)D(3) and increases steady state mRNA levels of both OPN and osteocalcin over a 48 h period. Together, these two analogs recapitulate all of the known actions of 1alpha,25(OH)(2)D(3) on osteoblasts. Based on these findings, we conclude that Ca(2+) release from bone stimulated by 1alpha,25(OH)(2)D(3) and PTH is related to the rapid, membrane-initiated actions and is not likely to involve binding to the nuclear receptor for 1alpha,25(OH)(2)D(3). Longer term stimulation of bone formation by 1alpha,25(OH)(2)D(3), however, appears to involve solely the nuclear receptor-mediated effects. These findings support our model of 1alpha,25(OH)(2)D(3) as a coupling factor for bone resorption and formation during bone remodeling.  相似文献   

11.
The rapid, nongenomic effects of 1alpha,25-dihydroxyvitamin D3 (1alpha,25-(OH)2D3 have been related to a 1,25D3-membrane associated, rapid response steroid binding protein or 1,25D3-[MARRS]bp, with a molecular weight of 65 kDa, in several tissues and species. Currently, no information is available concerning the nongenomic responses to 1alpha,25-(OH)2D3 in dental tissues. In order to investigate the expression of 1,25D3-[MARRS]bp in dental cells, in the presence or absence of 1alpha,25-(OH)2D3, we have used rabbit polyclonal antibodies directed against the N-terminus of the 1,25D3-[MARRS]bp (Ab099) that recognizes the 1alpha,25-(OH)2D3 binding protein in chick intestinal basolateral membranes and a mouse odontoblast-like cell line (MO6-G3). Western blotting and flow cytometric analyses with Ab099 specifically detected 1,25D3-[MARRS]bp in MO6-G3 cells. Moreover, 1,25D3-[MARRS]bp was up-regulated, in vivo, in differentiated dental cells. Electron microscopic analysis confirmed the plasma membrane localization of this binding protein and also showed its intracellular presence. Incubation of MO6-G3 cells with different doses of 1alpha,25-(OH)2D3 for 36 h resulted in an inhibition of 1,25D3-[MARRS]bp expression with a maximal effect at 50 nM steroid. In addition, the culture media of MO6-G3 cells contains immunoreactive 1,25D3-[MARRS]bp. Immunogold positive membrane vesicle-like structures are present in the extracellular matrix of MO6-G3 cells. Altogether, these results indicate that the 1,25D3-[MARRS]bp expression in MO6-G3 cells is modulated by 1alpha,25-(OH)2D3. In conclusion, this 1alpha,25-(OH)2D3 binding protein could play an important role in the rapid, nongenomic responses to 1alpha,25-(OH)2D3 in dental cells.  相似文献   

12.
13.
14.
15.
Previous work from our laboratory demonstrated that 1,25(OH)2D3 rapidly stimulated hydrolysis of membrane polyphosphoinositides (PI) in rat colonocytes and in Caco-2 cells, generating the second messengers DAG and IP3. [Ca2+]i subsequently increased due to IP3-mediated release of intracellular Ca2+ stores, and to Ca2+ influx through a receptor-mediated Ca channel. Studies examining purified antipodal plasma membranes and experiments using Caco-2 cell monolayers found that 1,25(OH)2D3 influenced PI turnover only in the basolateral (BLM) and not brush border (BBM) membranes. Vitamin D analogues with poor affinity for the vitamin D receptor were found to effectively stimulate PI turnover, suggesting the presence of a unique vitamin D receptor in the BLM. Studies from our laboratory have demonstrated saturable, reversible binding of 1,25(OH)2 D3 to colonocyte BLM. Recently, we found that 1,25(OH)2D3 activated the tyrosine kinase c-src in colonocyte BLM by a heterotrimeric guanine nucleotide binding protein (G-protein)-dependent mechanism, with subsequent phosphorylation, translocation to the BLM, and activation of PI-specific phospholipase C gamma. Due to the rise in [Ca2+]i and DAG, two isoforms of protein kinase C (PKCalpha and PKCbeta2), but not other isoforms were activated by 1,25(OH)2D3 in rat colonocytes. Recent studies demonstrated that the seco-steroid translocated the beta2 isoform to the BLM, but not the BBM. In contrast, the alpha isoform did not translocate to either antipodal plasma membrane, but modulated IP3-mediated Ca2+ release from the endoplasmic reticulum. Preliminary studies have shown that 1,25(OH)2D3 also activated phosphatidylcholine phospholipase D (PLD) in Caco-2 cells, generating phosphatidic acid and contributing to the sustained rise in DAG. PLD stimulation occurred by both PKC-dependent and -independent mechanisms. Inhibitors of G-proteins, c-src, and PKC blunted the seco-steroid-mediated activation of PLD. Cells stably transfected with sense PKCalpha showed increased 1,25(OH)2D3-stimulated PLD activation, whereas transfectants with antisense PKCalpha had an attenuated response. In addition, 1,25(OH)2D3 also regulated PLD by activating the monomeric G-protein rho A by a mechanism independent of the G-protein/ c-src/PKC pathway.  相似文献   

16.
17.
Protein kinase C (PKC) is a signal transduction protein that has been proposed to mediate rapid responses to steroid hormones. Previously, we have shown aldosterone directly activates PKCalpha whereas 17beta-estradiol activates PKCalpha and PKCdelta; however, neither the binding to PKCs nor the mechanism of action has been established. To determine the domains of PKCalpha and PKCdelta involved in binding of aldosterone and 17beta-estradiol, glutathione S-transferase fusion recombinant PKCalpha and PKCdelta mutants were used to perform in vitro binding assays with [(3)H]aldosterone and [(3)H]17beta-estradiol. 17beta-Estradiol bound both PKCalpha and PKCdelta but failed to bind PKC mutants lacking a C2 domain. Similarly, aldosterone bound only PKCalpha and mutants containing C2 domains. Thus, the C2 domain is critical for binding of these hormones. Binding affinities for aldosterone and 17beta-estradiol were between 0.5-1.0 nM. Aldosterone and 17beta-estradiol competed for binding to PKCalpha, suggesting they share the same binding site. Phorbol 12,13-dybutyrate did not compete with hormone binding; furthermore, they have an additive effect on PKC activity. EC(50) for activation of PKCalpha and PKCdelta by aldosterone and 17beta-estradiol was approximately 0.5 nM. Immunoblot analysis using a phospho-PKC antibody revealed that upon binding, PKCalpha and PKCdelta undergo autophosphorylation with an EC(50) in the 0.5-1.0 nm range. 17beta-Estradiol activated PKCalpha and PKCdelta in estrogen receptor-positive and -negative breast cancer cells (MCF-7 and HCC-38, respectively), suggesting estrogen receptor expression is not required for 17beta-estradiol-induced PKC activation. The present results provide first evidence for direct binding and activation of PKCalpha and PKCdelta by steroid hormones and the molecular mechanisms involved.  相似文献   

18.
Pretreatment of freshly isolated human peripheral blood monocytes with the steroid hormone, 1 alpha,25-dihydroxyvitamin D3 (1,25(OH)D), markedly reduced (by 95%) productive infection of human monocytes by HIV-1. Equivalent concentrations (10nM) of 25-hydroxyvitamin D3 (25(OH)D), the biologic precursor of 1,25(OH)D, were ineffective at reducing either CD4 expression or HIV-1 production. Pretreatment was required for modulation of HIV-1 infection by 1,25(OH)D. Interestingly, 1,25(OH)D-mediated decreases in p24 antigen production were observed prior to any observed reduction in CD4 expression, suggesting that 1,25(OH)D treatment may modulate HIV-1 infection of monocytes through additional factors besides decreased HIV-1 binding. These data raise the possibility that 1,25(OH)D compounds may be important in host resistance to HIV-1.  相似文献   

19.
Previous studies have indicated that upon in vitro activation with mitogenic lectins, human peripheral blood T lymphocytes express receptors for the steroid hormone 1 alpha, 25-dihydroxyvitamin D3(1,25(OH)2D3). Furthermore, the hormone can inhibit interleukin 2 production by the activated cells. In this investigation, we report that human peripheral B lymphocytes activated in vitro with the B lymphotropic Epstein-Barr virus (EBV) also express 1,25(OH)2D3 receptor-like macromolecules. These receptors are localized in the cell nucleus and exhibit properties similar to those found in classical target tissues for 1,25(OH)2D3. They sediment on sucrose gradients at 3.3 S, display a dissociation constant (Kd) of 4 X 10(-10) M, and can bind to DNA. In addition to the 1,25(OH)2D3 receptors, however, EBV-activated lymphocytes express a second class of 1,25(OH)2D3-binding proteins that appear to occur mainly in the cell cytosol and exhibit distinct biochemical properties from the receptor, including higher sedimentation coefficients (3.7 S to 4 S) and the lack of ability to bind to DNA. The addition of 1,25(OH)2D3 to cultures of EBV-infected cells inhibited the production of IgM and IgG by the B cells. The vitamin D3 analog 24,25(OH)2D3 did not inhibit Ig production, thus suggesting that the effect is probably mediated through the high affinity receptor macromolecule localized in the nucleus. Because the EBV-induced Ig production is independent of T cell participation, the data also suggest that the effects of 1,25(OH)2D3 are exerted directly on the B cell. The present results add to the evidence of the importance of 1,25(OH)2D3 as an immunoregulatory hormone.  相似文献   

20.
Synthesis of type I and III collagens has been examined in MG-63 human osteosarcoma cells after treatment with the steroid hormone, 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3). Analysis of total [3H]proline-labeled proteins and pepsin-derived collagens revealed that 1,25-(OH)2D3 selectively stimulated synthesis of alpha 1I and alpha 2I components of type I collagen after 6-12 h. Consistent with previous reports (Franceschi, R. T., Linson, C. J., Peter, T. C., and Romano, P. R. (1987) J. Biol. Chem. 262, 4165-4171), parallel increases in fibronectin synthesis were also observed. Hormonal effects were maximal (2- to 2.5-fold versus controls) after 24 h and persisted for at least 48 h. In contrast, synthesis of the alpha 1III component of type III collagen was not appreciably affected by hormone treatment. Of several vitamin D metabolites (1,25-(OH)2D3, 25-dihydroxyvitamin D3, and 24R,25-dihydroxyvitamin D3) tested for activity in stimulating type I collagen synthesis, 1,25-(OH)2D3 was found to be the most active. Analysis of collagen mRNA abundance by Northern blot hybridization indicated that both types I and III procollagen mRNAs were increased 4-fold after a 24-h exposure to 1,25-(OH)2D3. Pro alpha 1I mRNA remained elevated through the 48-h time point while pro alpha 2I and pro alpha 1III mRNAs returned to control values. These results indicate that the regulation of collagen synthesis by 1,25-(OH)2D3 is complex and may involve changes in translational efficiency as well as mRNA abundance. 1,25-(OH)2D3 also caused at least a 20-fold increase in levels of the bone-specific calcium-binding protein, osteocalcin. These results are consistent with the hypothesis that 1,25-(OH)2D3 is stimulating partial differentiation to the osteoblast phenotype in MG-63 cells.  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