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 0 毫秒
1.
Vitamin D target proteins: function and regulation   总被引:13,自引:0,他引:13  
  相似文献   

2.
Activation of precursor 25‐hydroxyvitamin D3 (25D) to hormonal 1,25‐dihydroxyvitamin D3 (1,25D) is a pivotal step in vitamin D physiology, catalysed by the enzyme 25‐hydroxyvitamin D‐1α‐hydroxylase (1α‐hydroxylase). To establish new models for assessing the physiological importance of the 1α‐hydroxylase‐25D‐axis, we used Danio rerio (zebrafish) to characterize expression and biological activity of the gene for 1α‐hydroxylase (cyp27b1). Treatment of day 5 zebrafish larvae with inactive 25D (5–150 nM) or active 1,25D (0.1–10 nM) induced dose responsive expression (15–95‐fold) of the vitamin D‐target gene cyp24a1 relative to larvae treated with vehicle, suggesting the presence of Cyp27b1 activity. A full‐length zebrafish cyp27b1 cDNA was then generated using RACE and RT‐PCR methods. Sequencing of the resulting clone revealed an open reading frame encoding a protein of 505 amino acids with 54% identity to human CYP27B1. Transfection of a cyp27b1 expression vector into HKC‐8, a human kidney proximal tubular epithelial cell line, enhanced intracrine metabolism of 25D to 1,25D resulting in greater than twofold induction of CYP24A1 mRNA expression and a 25‐fold increase in 1,25D production compared to empty vector. These data indicate that we have cloned a functional zebrafish CYP27B1, representing a phylogenetically distant branch from mammals of this key enzyme in vitamin D metabolism. Further analysis of cyp27b1 expression and activity in zebrafish may provide new perspectives on the biological importance of 25D metabolism. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

3.
4.
The Steroid hormon 1α, @5-Dihydroxyvitamin D3 has been shown to expert rapid effect (15 s to 5 min) in osteoblast. These occur in osteoblast-like cells lacking the nuclear vitamin D receptor, ROS 24/1, suggesting that a separate signalling system mediates the rapid action. These non-genomic action include rapid activation of phospholipase C and opening of calcium channels, pointing to a membrane localization of this signalling system. Previous studies have shown that the 1β epimer of 1α25-dihydroxyvitamina D3 can block these rapid action, indicating that the 1β epimer may bind to the recptor responsible for the rapid action sin a competative manner. We have assessed the displacement of 3H-1α,25dihydroxyvitamin D3 by vitamin D compounds, as well as the apparent dissociation constant of 1α25-dihydroxyvitamin D3 and its 1β epimer for the memberane receptor in membrane prepration from ROS 24/1 cells. Increasing concentrations of 1α25-dihydroxyvitamin D3, 7.25 nM to 725 nM, displaced 3H-1α25-dihydrxyvitamin D3 from the membranes with 725 nM of the hormone displacing 40–49% of the radioactivity. Similarly, 1β,25-dihydroxyvitamin D3, 7.25 nM and 72.5 nM, displaced 1α25-dihydroxyvitamin D3 binding while 25-hydroxyvitamin D3, 7.25 nM, did not. The apparent dissociation constant (KD) for 1α25-dihydroxyvitamin D3 was detrermined from displacement of 3H-1α25-dihydroxyvitamin D3 yielding a value of 8.1 × 10?7 M by Scatchard analysis. The KD for the 1β epimer determine from displacement of 3H-1α25-dihydroxyvitamin D3 was 4.8 × 10?7 M. The data suggest the presence of a receptor on the membrane of ROS 24/1 cells that reconize 1α25-dihydroxyvitamin D3 and its 1β epimer, but not 25-dihydroxyvitamin D3. Its ability to reconize the 1β epimer which appears to be a specific anagonist of the rapid effect of the hormone suggests that these studies may be the initial steps in the isolation and characterization of the signalling system mediating the rapid action of vitamin D.  相似文献   

5.
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.
New insights into the mechanisms of vitamin D action   总被引:17,自引:0,他引:17  
  相似文献   

8.
9.
10.
The seco-steroid hormone 1alpha,25(OH)(2)-vitamin D(3) (1,25-D(3)) is known to generate biological responses via both genomic and non-genomic rapid signal transduction pathways. The calcium regulated annexin II/p11 heterotetramer (AII(2)/p11(2)] was proposed by Baran and co-authors to be the membrane receptor responsible for mediating non-genomic, rapid actions of 1,25-D(3), based on ligand affinity labeling, competition, and saturation analysis experiments. Given the cytosolic presence of both the monomeric and heterotetrameric form of AII and their functional regulation by intracellular calcium concentrations, which are known to be affected by 1,25-D(3) rapid, non-genomic activities, we investigated in vitro the affinity of [(3)H]1,25-D(3) for the AII monomer and AII(2)/p11(2) in the absence and presence of calcium using saturation analysis and gel-filtration chromatography. Using two different techniques for separating bound from free ligand (perchlorate and hydroxylapatite (HAP)) over a series of 30 experiments, no evidence for specific binding of [(3)H]1,25-D(3) was obtained with or without the presence of 700 nM exogenous calcium, using either the AII monomer or AII(2)/p11(2). However saturable binding of [(3)H]1,25-D(3) to the lipid raft/caveolae enriched rat intestinal fraction was consistently observed (K(d) = 3.0 nM; B(max) = 45 fmols/mg total protein). AII was detected in lipid raft/caveolae enriched fractions from rat and mouse intestine and ROS 17/2.8 and NB4 cells by Western blot, but incubation in the presence of exogenous calcium did not ablate 1,25-D(3) binding as reported by Baran et al. Our results suggest that AII does not bind 1,25-D(3) in a physiologically relevant manner; however, recent studies linking AII(2)/p11(2) phosphorylation to vesicle fusion and its calcium regulated localization may make AII a possible down-stream substrate for 1,25-D(3) induced rapid cellular effects.  相似文献   

11.
12.
The receptors for retinoic acid (RA) and for 1α,25-dihydroxyvitamin D3 (VD), RAR, RXR, and VDR are ligand-inducible members of the nuclear receptor superfamily. These receptors mediate their regulatory effects by binding as dimeric complexes to response elements located in regulatory regions of hormone target genes. Sequence scanning of the tumor necrosis factor-α type I receptor (TNFαRI) gene identified a 3′ enhancer region composed of two directly repeated hexameric core motifs spaced by 2 nucleotides (DR2). On this novel DR2-type sequence, but not on a DR5-type RA response element, VD was shown to act through its receptor, the vitamin D receptor (VDR), as a repressor of retinoid signalling. The repression appears to be mediated by competitive protein–protein interactions between VDR, RAR, RXR, and possibly their cofactors. This VDR-mediated transrepression of retinoid signaling suggests a novel mechanism for the complex regulatory interaction between retinoids and VD. J. Cell. Biochem. 67:287–296, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

13.
David E. Kenny 《Zoo biology》2007,26(3):167-173
After the development of rickets in two hand‐reared polar bear (Ursus maritimus) cubs at the Denver Zoological Foundation we initiated a project opportunistically evaluating sera 25‐OH‐D3, calcium, phosphorus, and alkaline phosphatase levels in seven mother‐reared polar bear cubs. From one to three serum samples were obtained while the cub was still in the den with no access to natural sunlight and one after the cub exited the den. The sera 25‐OH‐D3 increased as the cubs grew (108±37 nmol/L [n=7] while in den, 184.6±48.3 nmol/L [n=7] out of the den) but were substantially lower then captive adults (348±215 nmol/L [n=36]) or free‐ranging adults (360±135 nmol/L [n=56]). We found that age was correlated significantly with 25‐OH‐D3 levels but den status (in or out) was not significant. Sera calcium and phosphorus values were highest in denned cubs (calcium, 2.9±0.4 mmol/L [n=13]; phosphorus, 3.1±0.5 mmol/L [n=13]) and then declined as the cubs matured and left the den (calcium, 2.5±0.1 mmol/L [n=7]; phosphorus, 2.4±0.4 mmol/L [n=7]). The enzyme alkaline phosphatase (presumably the bone isoenzyme) was high in denned growing cubs (831±426 IU/L [n=13]) and then declined gradually as the cubs aged and exited the den (96±83 IU/L [n=7]). Zoo Biol 0:1–7, 2007. © 2007 Wiley‐Liss, Inc.  相似文献   

14.
Evidence suggests that environmental and dietary factors may contribute to the pathogenesis of Parkinson’s disease (PD). High dietary intake of cholesterol is such a factor that has been shown to increase or decrease the risk of PD. However, because circulating cholesterol does not cross the blood–brain barrier, the mechanisms linking dietary cholesterol to the pathogenesis of PD remain to be understood. In contrast to cholesterol, the oxidized cholesterol metabolites (oxysterols), 24S‐hydroxycholesterol (24‐OHC) and 27‐hydroxycholesterol (27‐OHC), can cross the blood–brain barrier and may place the brain at risk of degeneration. In this study, we incubated the human neuroblastoma SH‐SY5Y cells for 24 h with 24‐OHC, 27‐OHC, or a mixture of 24‐OHC plus 27‐OHC, and have determined effects on tyrosine hydroxylase (the rate‐limiting enzyme in dopamine synthesis) levels, α‐synuclein levels, and apoptosis. We demonstrate that while 24‐OHC increases the levels of tyrosine hydroxylase, 27‐OHC increases levels of α‐synuclein, and induces apoptosis. Our findings show for the first time that oxysterols trigger changes in levels of proteins that are associated with the pathogenesis of PD. As steady state levels of 24‐OHC and 27‐OHC are tightly regulated in the brain, disturbances in these levels may contribute to the pathogenesis of PD.  相似文献   

15.
16.
The actions of the hormonal form of vitamin D, 1α,25-dihydroxyvitamin D3 [1α,25-(OH)2D3], are mediated by both genomic and nongenomic mechanisms. Several vitamin D synthetic analogs have been developed in order to identify and characterize the site(s) of action of 1α,25-(OH)2D3 in many cell types including osteoblastic cells. We have compared the effects of 1α,25-(OH)2D3 and a novel 1α,25-(OH)2D3 bromoester analog (1,25-(OH)2-BE) that covalently binds to vitamin D receptors. Rat osteosarcoma cells that possess (ROS 17/2.8) or lack (ROS 24/1) the classic intracellular vitamin D receptor were studied to investigate genomic and nongenomic actions. In ROS 17/2.8 cells plated at low density, the two vitamin D compounds (1 × 10−8 M) caused increased cell proliferation, as assessed by DNA synthesis and total cell counts. Northern blot analysis revealed that the mitogenic effect of both agents was accompanied by an increase in steady-state osteocalcin mRNA levels, but neither agent altered alkaline phosphatase mRNA levels in ROS 17/2.8 cells. ROS 17/2.8 cells responded to 1,25-(OH)2-BE but not the natural ligand with a significant increase in osteocalcin secretion after 72, 96, 120, and 144 hr of treatment. Treatment of ROS 17/2.8 cells with the bromoester analog also resulted in a significant decrease in alkaline phosphatase-specific activity. To compare the nongenomic effects of 1α,25-(OH)2D3 and 1,25-(OH)2-BE, intracellular calcium was measured in ROS 24/1 cells loaded with the fluorescent calcium indicator Quin 2. At 2 × 10−8 M, both 1α,25-(OH)2D3 and 1,25-(OH)2-BE increased intracellular calcium within 5 min. Both the genomic and nongenomic actions of 1,25-(OH)2-BE are similar to those of 1α,25-(OH)2D3, and since 1,25-(OH)2-BE has more potent effects on osteoblast function than the naturally occurring ligand due to more stable binding, this novel vitamin D analog may be useful in elucidating the structure and function of cellular vitamin D receptors. © 1996 Wiley-Liss, Inc.  相似文献   

17.
The vitamin D3 receptor (VDR) is present in all microenvironments of the breast, yet it is hypothesized to signal through the epithelium to regulate hormone induced growth and differentiation. However, the influence or contribution of the other microenvironments within the breast that express VDR, like the breast adipose tissue, are yet to be investigated. We hypothesized that the breast adipocytes express the signaling components necessary to participate in vitamin D3 synthesis and signaling via VDR, modulating ductal epithelial cell growth and differentiation. We utilized human primary breast adipocytes and VDR wild type (WT) and knockout (KO) mice to address whether breast adipocytes participate in vitamin D3‐induced growth regulation of the ductal epithelium. We report in this study that breast primary adipocytes express VDR, CYP27B1 (1α‐hydroxylase, 1α‐OHase), the enzyme that generates the biologically active VDR ligand, 1α,25‐dihydroxyvitamin D3 (1,25D3), and CYP24 (24‐hydroxylase, 24‐OHase), a VDR‐1,25D3 induced target gene. Furthermore, the breast adipocytes participate in bioactivating 25‐hydroxyvitamin D3 (25D3) to the active ligand, 1,25D3, and secreting it to the surrounding microenvironment. In support of this concept, we report that purified mammary ductal epithelial fragments (organoids) from VDR KO mice, co‐cultured with WT breast adipocytes, were growth inhibited upon treatment with 25D3 or 1,25D3 compared to vehicle alone. Collectively, these results demonstrate that breast adipocytes bioactivate 25D3 to 1,25D3, signal via VDR within the adipocytes, and release an inhibitory factor that regulates ductal epithelial cell growth, suggesting that breast adipose tissue contributes to vitamin D3‐induced growth regulation of ductal epithelium. J. Cell. Biochem. 112: 3393–3405, 2011. © 2011 Wiley Periodicals, Inc.  相似文献   

18.
19.
1,25(OH)2D3 is an antiproliferative agent that may inhibit proliferation of breast cancer (BC) cells in vitro and BC development in animals. Epidemiological studies have shown a high incidence of BC in people less exposed to solar rays. To unravel the role of Vitamin D3 in BC patients, we have investigated serum levels of 25(OH)D3 and its active form 1,25(OH)2D3 as well as tissue expression of 1alpha-hydroxylase, 24-hydroxylase, and Vitamin D-receptor (VDR), determined by semiquantitative RT-PCR, in 88 Brazilian BC patients and 35 women without cancer (submitted to mammoplasties or resection of benign lesions). Median age of women with and without cancer was 51 and 46 years, respectively, and the majority of BC patients were classified as clinical stage II (67%). Although no differences in 25(OH)D3 serum concentration were found, 1,25(OH)2D3 (40+/-21 pg/ml) levels in BC patients were lower than in women without cancer (53+/-23). Our results indicate that 24-hydroxylase, VDR and 1alpha-hydroxylase mRNA tissue expression is similar in both groups and no correlation between 24-hydroxylase, 1alpha-hydroxylase, and VDR expression in breast tumors was found. A low 1,25(OH)2D3 serum concentration seems to be associated to breast cancer, however, the mechanism involved in this regulation is still unclear.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号