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 31 毫秒
1.
We have recently reported that annexin II serves as a membrane receptor for 1α,25‐(OH)2D3 and mediates the rapid effect of the hormone on intracellular calcium. The purpose of these studies was to characterize the binding of the hormone to annexin II, determine the specificity of binding, and assess the effect of calcium on binding. The binding of [14C]‐1α,25‐(OH)2D3 bromoacetate to purified annexin II was inhibited by 1α,25‐(OH)2D3 in a concentration‐dependent manner. Binding of the radiolabeled ligand to annexin II was markedly diminished by 1α,25‐(OH)2D3 at 24 μM, 18 μM, and 12 μM and blunted by 6 μM and 3 μM. At a concentration of 12 μM, 1β,25‐(OH)2D3 also diminished the binding of [14C]‐1α,25‐(OH)2D3 bromoacetate to annexin II, but cholecalciferol, 25‐(OH)D3, and 24,25‐(OH)2D3 did not. Saturation analyses of the binding of [3H]‐1α,25‐(OH)2D3 to purified annexin II showed a KD of 5.5 × 10−9 M, whereas [3H]‐1β,25‐(OH)2D3 exhibited a KD of 6.0 × 10−9 M. Calcium, which binds to the carboxy terminal domain of annexin II, had a concentration‐dependent effect on [14C]‐1α,25‐(OH)2D3 bromoacetate binding to annexin II, with 600 nM calcium being able to inhibit binding of the radiolabeled analog. The inhibitory effect of calcium was prevented by EDTA. Homocysteine, which binds to the amino terminal domain of annexin II, had no effect on the binding of the bromoacetate analog to the protein. The data indicate that 1α,25‐(OH)2D3 binding to annexin II is specific and suggest that the binding site may be located on the carboxy terminal domain of the protein. The ability of 1β,25‐(OH)2D3 to inhibit the binding of [14C]‐1α,25(OH)2D3 bromoacetate to annexin II provides a biochemical explanation for the ability of the 1β‐epimer to inhibit the rapid actions of the hormone in vitro. J. Cell. Biochem. 80:259–265, 2000. © 2000 Wiley‐Liss, Inc.  相似文献   

2.
BackgroundEpidemiological studies indicate high serum 25(OH)D3 is associated with increased survival in breast cancer patients. Pre-clinical studies attributed this to anti-tumorigenic properties of its metabolite 1α,25(OH)2D3. However, 1α,25(OH)2D3 is highly calcemic and thus has a narrow therapeutic window. Here we propose another metabolite, 24R,25(OH)2D3, as an alternative non-calcemic vitamin D3 supplement.MethodsNOD-SCID-IL2γR null female mice with MCF7 breast cancer xenografts in the mammary fat pad were treated with 24R,25(OH)2D3 and changes in tumor burden and metastases were assessed. ERα66+ MCF7 and T47D cells, and ERα66- HCC38 cells were treated with 24R,25(OH)2D3 in vitro to assess effects on proliferation and apoptosis. Effects on migration and metastatic markers were assessed in MCF7.Results24R,25(OH)2D3 reduced MCF7 tumor growth and metastasis in vivo. In vitro results indicate that this was not due to an anti-proliferative effect; 24R,25(OH)2D3 stimulated DNA synthesis in MCF7 and T47D. In contrast, markers of invasion and metastasis were decreased. 24R,25(OH)2D3 caused dose-dependent increases in apoptosis in MCF7 and T47D, but not HCC38 cells. Inhibitors to palmitoylation, caveolae integrity, phospholipase-D, and estrogen receptors (ER) demonstrate that 24R,25(OH)2D3 acts on MCF7 cells through caveolae-associated, phospholipase D-dependent mechanisms via cross-talk with ERs.ConclusionThese results indicate that 24R,25(OH)2D3 shows promise in treatment of breast cancer by stimulating tumor apoptosis and reducing metastasis.General significance24R,25(OH)2D3 regulates breast cancer cell survival through ER-associated mechanisms similar to 24R,25(OH)2D3 effects on chondrocytes. Thus, 24R,25(OH)2D3 may modulate cell survival in other estrogen-responsive cell types, and its therapeutic potential should be investigated in ER-associated pathologies.  相似文献   

3.
The effect of vitamin D metabolites on the growth of chick embryo chondrocytes in soft agar was examined. 1,25-Dihydroxyvitamin D3 [1,25(OH)2D3]at 10−8-10−7 M induced colony formation by chick embryo chondrocytes in soft agar in the presence of 10% fetal bovine serum. Furthermore, 1,25(OH)2D3 increased the number of colonies in the presence of a maximal dose of basic fibroblast growth factor, a potent mitogen for chondrocytes in soft agar. However, 24R,25 (OH)2D3 and other metabolites had little effect on the soft agar growth of chondrocytes in the presence or absence of basic fibroblast growth factor. These results suggest that 1,25(OH)2D3 is an active metabolite which may be involved in supporting cartilage growth.  相似文献   

4.
The essential role of vitamin D throughout the life of most mammals and birds as a mediator of calcium homeostasis is well established. In view of the complex endocrine system existent for the regulated metabolism of vitamin D3 to both 1α,25-dihydroxyvitamin D3 [1,25(OH)2D3] and 24R,25-dihydroxyvitamin D3 [24R,25-(OH)2D3] (both produced by the kidney), an intriguing problem is to elucidate whether only one or both of these dihydroxyvitamin D3 metabolites is required for the generation of all the biological responses mediated by the parent vitamin D3. In contrast to the accumulated knowledge concerning the short term actions of 1,25(OH)2-D3 on stimulating intestinal calcium absorption and bone calcium reabsorption, relatively little is known of the biological function of 24,25(OH)2D3. We report now the results of a nine month study in which chicks were raised on a vitamin D-deficient diet from hatching to sexual maturity and received as their sole source of “vitamin D” either 24,25(OH)2D3 or 1,25(OH)2D3 singly or in combination. Specifically we are describing the integrated operation of the vitamin D endocrine system as quantitated by the individual measurement in all birds of 22 variables related to “vitamin D status” and as evaluated by the statistical procedure of multivariate discriminant analysis. Twelve of these variables involved detailed analysis of the bone including quantitative histology and the other 10 variables reflect various manifestations of vitamin D action, e.g. serum Ca2+ and Pi levels, vitamin D-dependent calcium binding protein (CaBP) in the intestine and kidney, egg productivity etc. As evaluated by the multivariate analysis, it is clear that 24,25(OH)2D3 and 1,25(OH)2D3 are simultaneously required for normalization of calcium homeostasis.  相似文献   

5.
The effects of treatment with the osteotropic steroids 1,25-dihydroxyvitamin D3 (1,25(OH)2D3), 17β-estradiol, or dexamethasone on [1-14C]arachidonic acid (AA) uptake and distribution into glycerophospholipid classes by normal adult human osteoblast-like (hOB) cells were investigated. Total uptake of [1-14C]AA was decreased in cells treated with dexamethasone when assayed after a 24-, 48-, or 96-h exposure to the hormone. Specific radiolabel incorporation into phosphatidylcholine was reduced by a 48-h treatment with dexamethasone with a concurrent increase in the radiolabeling of phosphatidylethanolamine. However, these changes were transient, and by 96 h of dexamethasone treatment the distribution of the radiolabeled fatty acid had reequilibrated to resemble the pattern found for vehicle treated samples. Total uptake of [1-14C]AA was diminished by 96-h treatment with 1,25(OH)2D3 (79 ± 3% of control, P < 0.01); at that time point, a significant decrease in the proportional radiolabeling of the phosphatidylinositol pool was identified (92 ± 2% of control, P < 0.05). The 1,25(OH)2D3-dependent decrease in total uptake and in phosphatidylinositol incorporation of [1-14C]AA were found to be hormone dose dependent. Treatment with 24,25(OH)2D3 was without effect on either total [1-14C]AA uptake or the specific [1-14C]AA radiolabeling of the phosphatidylinositol pool. 1,25(OH)2D3 treatment decreased hOB cell uptake of [1-14C]oleic acid and decreased its proportional incorporation into the phosphatidylinositol pool. Gas chromatographic analyses revealed no 1,25(OH)2D3-dependent effects on total phosphatidylinositol lipid mass or on the mole percent of arachidonic acid within the phosphatidylinositol pool, leaving the mechanism of the effects of the secosteroid on hOB cell AA metabolism unexplained. 17β-Estradiol had no effects on the parameters of AA metabolism measured. As a consequence of their modulation of arachidonic acid uptake and its distribution into hOB cellular phospholipids, steroids might alter the biological effects of other hormones whose actions include the stimulated production of bioactive AA metabolites, such as prostaglandins or the various lipoxygenase products.  相似文献   

6.
Stumpf  W. E.  Koike  N.  Hayakawa  N.  Tokuda  K.  Nishimiya  K.  Tsuchiya  Y.  Hirate  J.  Okazaki  A.  Kumaki  K. 《Histochemistry and cell biology》1994,102(3):183-194
Target cells for 3H-labeled 1, 25(OH)2 vitamin D3 [1,25(OH)2D3, vitamin D] and its analog 3H-labeled 22-oxa-1, 25(OH)2 vitamin D3 (OCT) have been identified during endochondral and intramembranous ossification in developing, undecalcified, unembedded bone, using thaw-mount autoradiography. Two-day-old neonatal rats were injected with [3H]1,25(OH)2D3 or [3H]OCT; after 2 h leg, spine, and head were frozen and sectioned. In the epiphyseal-metaphyseal region specific nuclear concentrations of [3H]1,25(OH)2D3 and [3H]OCT were observed in identical cell populations, being low in cells of the articular and resting zone, intermediate in the proliferating zone, and highest in hypertrophic chondrocytes and in osteoblasts and precursor cells. In the primary spongiosa intertrabecular spaces there were a large number of cells with nuclear labeling — probably osteoblasts and precursor cells. In contrast, in the secondary spongiosa intertrabecular spaces, apparent blood-forming cells were mostly unlabeled. Osteoblasts along bone spicules and compact bone in long bones, vertebrae, and head also showed strong nuclear labeling, as did cells of the periosteum. These data suggest that 1,25(OH)2D3 and OCT regulate development, differentiation, and activities of chondrocytes and osteoblasts, including differentiation of resting chondrocytes into proliferating and hypertrophic chondrocytes that involve chondroclastic enlargement of lacunae and trans-differentiation of surviving hypertrophic chondrocytes; differentiation of stroma cells into osteoblasts; and in periosteum and other regions of intramembranous ossification differentiation of precursor cells and osteoblasts. Nuclear receptor binding and their selective and hierarchical distribution during cell differentiation appear to correspond to multiple genomic effects toward growth, regeneration and repair. The findings indicate a physiological significance and therapeutic potential of 1,25(OH)2D3 and in particular of its less hypercalcemic analog OCT.On leave from the University of North Carolina  相似文献   

7.
8.
1,25-Dihydroxyvitamin D3 [1,25-(OH)2D3] receptor was characterized after partial purification of thymus cytosol by ammonium sulfate fractionation. The 1,25-(OH)2D3 receptor sediments at 3.7S in 5–20% sucrose gradients. The binding of 1,25-(OH)2D3 in thymic cytosol was a saturable process with high affinity (Kd = 0.12?0.48 nM) at 4°C. Competition for 1,25-(OH)2[3H]D3 receptor by nonradioactive analogs demonstrated the affinities of these analogs to be in order; 1,25-(OH)2D3 = 1,24R,25-(OH)3D3 = 1,25S,26-(OH)3D3 = 1,25R,26-(OH)3D3 > 1,25-(OH)2D3-26,23 lactone > 25-OHD3 > 23R,25-(OH)2D3 > 24R,25-(OH)2D3 > 23S,25-(OH)2D3 ? 25-OHD3-26,23 lactone. The receptor bound to DNA cellulose columns in low salt buffer and eluted as a single peak at 0.21 M KCl. These findings provide evidence that the thymus possesses a 1,25-(OH)2D3 receptor with properties indistinguishable from 1,25-(OH)2D3 receptors in other tissues.  相似文献   

9.
A new metabolite of vitamin D3 has been isolated in pure form from incubations of rat kidney homogenates with 25-hydroxyvitamin D3 [25-OH-D3]. It was identified as 23,25-dihydroxy-24-oxo-vitamin D3 [23,25(OH)2-24-oxo-D3] by means of ultraviolet absorption spectrophotometry and mass spectrometry. Also, 25-OH-D3-26,23-lactone and 24R,25-dihydroxyvitamin D3 were obtained from the same incubation mixtures. The enzyme activity responsible for the conversion of 25-OH-D3 to 23,25(OH)2-24-oxo-D3 was induced by perfusion of the kidneys invitro with 50 nM 1,25-dihydroxyvitamin D3 [1,25(OH)2D3].  相似文献   

10.
The endocrine feedback loop between vitamin D3 (1,25(OH)2D3) and parathyroid hormone (PTH) plays a central role in skeletal development. PTH‐related protein (PTHrP) shares homology and its receptor (PTHR1) with PTH. The aim of this study was to investigate whether there is a functional paracrine feedback loop between 1,25(OH)2D3 and PTHrP in the growth plate, in parallel with the endocrine feedback loop between 1,25(OH)2D3 and PTH. This was investigated in ATDC5 cells treated with 10?8 M 1,25(OH)2D3 or PTHrP, Col2‐pd2EGFP transgenic mice, and primary Col2‐pd2EGFP growth plate chondrocytes isolated by FACS, using RT‐qPCR, Western blot, PTHrP ELISA, chromatin immunoprecipitation (ChIP) assay, silencing of the 1,25(OH)2D3 receptor (VDR), immunofluorescent staining, immunohistochemistry, and histomorphometric analysis of the growth plate. The ChIP assay confirmed functional binding of the VDR to the PTHrP promoter, but not to the PTHR1 promoter. Treatment with 1,25(OH)2D3 decreased PTHrP protein production, an effect which was prevented by silencing of the VDR. Treatment with PTHrP significantly induced VDR production, but did not affect 1α‐ and 24‐hydroxylase expression. Hypertrophic differentiation was inhibited by PTHrP and 1,25(OH)2D3 treatment. Taken together, these findings indicate that there is a functional paracrine feedback loop between 1,25(OH)2D3 and PTHrP in the growth plate. 1,25(OH)2D3 decreases PTHrP production, while PTHrP increases chondrocyte sensitivity to 1,25(OH)2D3 by increasing VDR production. In light of the role of 1,25(OH)2D3 and PTHrP in modulating chondrocyte differentiation, 1,25(OH)2D3 in addition to PTHrP could potentially be used to prevent undesirable hypertrophic chondrocyte differentiation during cartilage repair or regeneration. J. Cell. Physiol. 229: 1999–2014, 2014. © 2014 Wiley Periodicals, Inc.
  相似文献   

11.
Uptake of orthophosphate (Pi) by osteoblast-like cells is known to be stimulated by parathyroid hormone (PTH), but effects on intracellular [Pi] have not been investigated. Here we show in rat osteoblast-like cells (UMR 106-06) that PTH (10?11 to 10?7 M ) increases both 32Pi uptake and cellular [Pi] by up to 50 per cent. 1,25 Dihydroxyvitamin D3 (1,25D) (10?12 to 10?6 M ) and salmon calcitonin (CT) (10?12 to 10?6 g ml?1) also increased cellular [Pi] (by up to 60 per cent), but the percentage increases in total cellular 32Pi uptake were smaller. The effects of 1,25D were transient (observable at 80 min and 6 h but not 24 h), and were also observed with 24,25 dihydroxy- and 25 hydroxyvitamin D3. Transient degradation of organic phosphorus pools to Pi might contribute to this increased [Pi]. These pools remain to be identified but were not shown to be phospholipids. Foetal bovine serum also affected cellular [Pi]. Care is therefore needed in distinguishing direct hormonal effects on cellular [Pi] from indirect effects arising from changes in the rate of cell growth.  相似文献   

12.
The most biologically active metabolite 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) has well known direct effects on osteoblast growth and differentiation in vitro. The precursor 25-hydroxyvitamin D3 (25(OH)D3) can affect osteoblast function via conversion to 1,25(OH)2D3, however, it is largely unknown whether 25(OH)D3 can affect primary osteoblast function on its own. Furthermore, 25(OH)D3 is not only converted to 1,25(OH)2D3, but also to 24R,25-dihydroxyvitamin D3 (24R,25(OH)2D3) which may have bioactivity as well. Therefore we used a primary human osteoblast model to examine whether 25(OH)D3 itself can affect osteoblast function using CYP27B1 silencing and to investigate whether 24R,25(OH)2D3 can affect osteoblast function. We showed that primary human osteoblasts responded to both 25(OH)D3 and 1,25(OH)2D3 by reducing their proliferation and enhancing their differentiation by the increase of alkaline phosphatase, osteocalcin and osteopontin expression. Osteoblasts expressed CYP27B1 and CYP24 and synthesized 1,25(OH)2D3 and 24R,25(OH)2D3 dose-dependently. Silencing of CYP27B1 resulted in a decline of 1,25(OH)2D3 synthesis, but we observed no significant differences in mRNA levels of differentiation markers in CYP27B1-silenced cells compared to control cells after treatment with 25(OH)D3. We demonstrated that 24R,25(OH)2D3 increased mRNA levels of alkaline phosphatase, osteocalcin and osteopontin. In addition, 24R,25(OH)2D3 strongly increased CYP24 mRNA. In conclusion, the vitamin D metabolites 25(OH)D3, 1,25(OH)2D3 and 24R,25(OH)2D3 can affect osteoblast differentiation directly or indirectly. We showed that primary human osteoblasts not only respond to 1,25(OH)2D3, but also to 24R,25(OH)2D3 by enhancing osteoblast differentiation. This suggests that 25(OH)D3 can affect osteoblast differentiation via conversion to the active metabolite 1,25(OH)2D3, but also via conversion to 24R,25(OH)2D3. Whether 25(OH)D3 has direct actions on osteoblast function needs further investigation.  相似文献   

13.
A competitive protein-binding radioassay for 24,25-dihydroxyvitamin D [24,25-(OH)2D] in human serum has been developed. Whereas small amounts of [3H]24,25-(OH)2D must be biosynthesized in order to trace the efficiency of the extraction and chromatographic procedures, tritiated 25-hydroxyvitamin D3 ([3H]25-OHD3) can be used as the assay tracer. Since 25-OHD3 and 24,25-(OH)2D3 are equipotent in their competitive displacement of [3H]25-OHD3 from rat serum, 25-OHD3 can be used as the assay standard. Liquid-gel partition chromatography on small columns of Sephadex LH-20 can reliably isolate 24,25-(OH)2D by batch elution. The purity of biosynthesized [3H]24,25-(OH)2D3 and the 24,25-(OH)2D fraction isolated from serum was confirmed by high-pressure chromatography on 0.2 × 50 cm columns of 10-μm silica. Serum 24,25-(OH)2D levels averaged 16% of the serum 25-OHD concentrations in normal subjects. Since chronic hemodialysis patients, without kidneys, had normal serum 24,25-(OH)2D levels, significant extrarenal 25-hydroxycalciferol 24-hydroxylase activity occurs in these subjects. Since the present assay represents a reasonably simple extension of 25-OHD assay methodology, it should prove to be a useful technique in the analysis of clinical disorders of vitamin D metabolism.  相似文献   

14.
The epidermis is both a target tissue for and a source of 1,25 dihydroxycholecal-ciferol. The present study determines which of the epidermal cell populations synthesizes 1,25 dihydroxycholecalciferol and which responds to this hormone. Epidermal keratinocytes from new born rat epidermis were separated by unit gravity sedimentation into poorly differentiated cells, slow-cycling more differentiated cells, actively proliferating cells, and terminally differentiating subpopulations. The keratinocyte populations were characterized by cell size analysis, cell morphology, and DNA and RNA contents (acridine orange flow cytometry). 1,25(OH)2D3 synthesis was studied by measuring the conversion of [3H] 25(OH)D3 to [3H] 1,25(OH)2D3. The purified product was tested for its ability to compete with synthetic [3H] 1,25(OH)2D3 for binding to chick intestinal cytosol. The responses of the keratinocyte subpopulations to exogeneous 1,25(OH)2D3 were evaluated by the increase in 25(OH)D3-24 hydroxylase activity. Furthermore the expression of 1,25(OH)2D3 receptors (VDR) was examined in these cell populations. The results show that only the least differentiated cells produced 1,25(OH)2D3. In contrast, immunocytochemical detection of VDR, the VDR mRNA, and a 25(OH)D3-24 hydroxylase response to 1,25(OH)2D3 were mainly found in the more differentiated cells. Thus, the ability of epidermis to synthesize 1,25(OH)2D3 and be simultaneously sensitive to it depends on the state of cell differentiation. This suggests that the mammalian epidermis contains a paracrine system in which the more differentiated keratinocytes are sensitive to the 1,25(OH)2D3 produced locally by neighboring immature ones. © 1994 wiley-Liss, Inc.  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.
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   

17.
Kidney homogenates from vitamin D3-supplemented chicks incubated with 25-hydroxyvitamin D3 [25(OH)D3] produce significant quantities of a new, unknown vitamin D metabolite. This metabolite was isolated in pure form from such incubation mixtures by using Sephadex LH-20 column chromatography followed by high-pressure liquid chromatography. This metabolite has been identified as 23,25,26-trihydroxyvitamin D3 [23,25,26(OH)3D3] by loss of radioactivity from 25-hydroxy[23,24-3H]vitamin D3 and 25-hydroxy-[26,27-methyl-3H]vitamin D3, ultraviolet absorption spectrophotometry, mass spectrometry, and periodate cleavage oxidation followed by mass spectrometry. This same metabolite was also isolated from the serum of rats given large doses of vitamin D3, and structurally characterized as 23,25,26-trihydroxyvitamin D3. As yet, the stereochemistry at the C-23 and C-25 positions of the natural product remains unknown. A comparison of responses to a single dose level (500 ng) of 23,25,26(OH)3D3 or 25(OH)D3 over 96 h in vitamin D-deficient rats indicated that the new metabolite had no capability to mediate bone calcium mobilization and that it was only weakly active in stimulating intestinal calcium transport.  相似文献   

18.
R and S isomers of 24-OH-D3 and 24,25-(OH)2D3 were tested for their effects on bone resorption in vitro. 24(R), 25-(OH)2D3 was more active than 24(S),25-(OH)2D3. Likewise, 24(R)-OH-D3 was more active than 24(S)-OH-D3. The bone resorbing activity of 24(R)-OH-D3 was equivalent to that of 25-OH-D3; 24(R),25-(OH)2D3 was somewhat less potent. The results indicate that there is discrimination between the isomers of these compounds at the level of the responding tissue.  相似文献   

19.
1,25-dihydroxyvitamin D3 (1,25-(OH)2D3) is known to be involved in regulating the proliferation of parathyroid cells and PTH synthesis through reactions involving its nuclear receptor. We evaluated the effects of 1,25-(OH)2D3 and its hexafluorinated analog, 26,26,26,27,27,27-hexafluoro-1,25-dihydroxyvitamin D3 (26,27-F6-1,25-(OH)2D3), on parathyroid cells. The 1,25-(OH)2D3 and 26,27-F6-1,25-(OH)2D3 each inhibited [3H]thymidine incorporation and ornithine decarboxylase (ODC) activity, which is important in cell proliferation, in primary cultured bovine parathyroid cells. The inhibitory effect of 26,27-F6-1,25-(OH)2D3 on PTH secretion from parathyroid cells was significantly more potent than that of 1,25-(OH)2D 3 between 10−11 M and 10−8 M. Study of 26,27-F6-1,25-(OH)2D3 metabolism in parathyroid cells in vitro elucidated its slower degradation than that of 1,25-(OH)2D3. After 48 h of incubation with [1β-3H]26,27-F6-1,25-(OH)2D3, two HPLC peaks, one for [1β-3H]26,27-F6-1,25-(OH)2D3, and a second larger peak for [1β-3H]26,27-F6-1,23(S),25-(OH)3D3, were detected. No metabolites were detected after the same period of incubation with 1,25-(OH)2[26,27-3H]D3. We observed that 26,27-F6-1,23(S),25-(OH)3D3 was as potent as 1,25-(OH)2D3 in inhibiting the proliferation of parathyroid cells.Data suggest that the greater biological activity of 26,27-F6-1,25-(OH)2D3 is explained by its slower metabolisms and by the retention of the biological potency of 26,27-F6-1,25-(OH)2D3 even after 23(S)-hydroxylation.  相似文献   

20.
Metabolism of 25-hydroxyvitamin D3 (25-OH-D3) in pregnancy was investigated invitro in New Zealand White rabbits fed a rabbit chow. Kidney homogenates from pregnant mothers and fetuses were separately incubated with [3H]-25-OH-D3. The homogenates from fetuses produced significant amounts of [3H]-1α,25-dihydroxyvitamin D3 [1α,25-(OH)2-D3] from its precursor, while those from mothers predominantly produced [3H]-24,25-dihydroxyvitamin D3 [24,25-(OH)2-D3]. The identity of the radioactive metabolites produced from [3H]-25-OH-D3 was established by periodate cleavage and comigration with synthetic 1α,25-(OH)2-D3 or 24,25-(OH)2-D3 on high pressure liquid chromatography. These results clearly indicate that the fetal kidney is at least one of the sites of 1α,25-(OH)2-D3 synthesis in pregnancy.  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