首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Vitamin D and androgen regulation of prostatic growth   总被引:1,自引:0,他引:1  
Vitamin D has been reported to inhibit the growth of prostate cancer cells and model systems. In this study, we examined the interaction between 1,25-dihydroxyvitamin D(3) (1,25 D) in the presence or absence of endogenous testosterone on the growth and development of the adult rat prostate. Male Sprague-Dawley rats (165 days old) were either kept intact or castrated. Seven days after castration, the rats were treated with vehicle (control) or 1,25 D for 3 weeks and then sacrificed. Both ventral and dorsal lateral prostates were harvested; whole tissue lysates were collected and AR and VDR protein levels were analyzed by immunoblot analyses. Administration of 1,25 D in the intact animals decreased the prostatic size by 40%, compared to control animals, whereas 1,25 D did not influence the size of the prostate in castrated rats. 1,25 D administration in intact groups also increased both the AR and VDR protein levels by approximately twofold, whereas in castrated groups, 1,25 D only increased the AR protein level by 1.5-2.5-fold. 1,25 D in the presence of endogenous testosterone inhibits prostatic growth, whereas 1,25 D in the absence of endogenous testosterone does not affect prostatic growth. The growth inhibitory activity of 1,25 D in the presence of testosterone may be mediated through the ligand activated AR and VDR pathways. These studies may reveal important information about the potential efficacy of 1,25 D as well as hormonal status in understanding the development of prostate diseases.  相似文献   

2.
Prostate cancer (PCa) cells express vitamin D receptors (VDR) and 1,25-dihydroxyvitamin D(3) (1,25(OH)(2)D(3)) inhibits the growth of epithelial cells derived from normal, benign prostate hyperplasia, and PCa as well as established PCa cell lines. The growth inhibitory effects of 1,25(OH)(2)D(3) in cell cultures are modulated tissue by the presence and activities of the enzymes 25-hydroxyvitamin D(3) 24-hydroxylase which initiates the inactivation of 1,25(OH)(2)D(3) and 25-hydroxyvitamin D(3) 1alpha-hydroxylase which catalyses its synthesis. In LNCaP human PCa cells 1,25(OH)(2)D(3) exerts antiproliferative activity predominantly by cell cycle arrest through the induction of IGF binding protein-3 (IGFBP-3) expression which in turn increases the levels of the cell cycle inhibitor p21 leading to growth arrest. cDNA microarray analyses of primary prostatic epithelial and PCa cells reveal that 1,25(OH)(2)D(3) regulates many target genes expanding the possible mechanisms of its anticancer activity and raising new potential therapeutic targets. Some of these target genes are involved in growth regulation, protection from oxidative stress, and cell-cell and cell-matrix interactions. A small clinical trial has shown that 1,25(OH)(2)D(3) can slow the rate of prostate specific antigen (PSA) rise in PCa patients demonstrating proof of concept that 1,25(OH)(2)D(3) exhibits therapeutic activity in men with PCa. Further investigation of the role of calcitriol and its analogs for the therapy or chemoprevention of PCa is currently being pursued.  相似文献   

3.
The aim of this study was to investigate effects of 1,25(OH)(2)D(3) (calcitriol), 25OHD(3), and EB1089 on cell growth and on Vitamin D receptor (VDR) mRNA and 1alpha-hydroxylase (1alpha-OHase) mRNA expression in normal canine prostatic primary cultures. Canine prostatic epithelial cells were isolated, cultured, and treated with vehicle (ethanol), calcitriol, 25OHD(3), and EB1089 at 10(-9) and 10(-7)M. The VDR was present in epithelial and stromal cells of the canine prostate gland. 1,25(OH)(2)D(3), 25OHD(3), and EB1089 inhibited epithelial cell growth at 10(-7)M compared to vehicle-treated controls [calcitriol (P < 0.01), EB1089 (P < 0.01), and 25OHD(3) (P < 0.05)]. Epithelial cells treated with calcitriol and EB1089 at 10(-7)M had slightly increased VDR mRNA expression (0.2-0.3-fold) at 6 and 12h compared to controls. There was no difference in 1alpha-OHase mRNA expression in epithelial cells treated with these three compounds. 1,25(OH)(2)D(3) and its analogs may be effective antiproliferative agents of epithelial cells in certain types of prostate cancer.  相似文献   

4.
The antitumor effects of 1,25-dihydroxyvitamin D3 (calcitriol) are being exploited for prevention and treatment of prostate cancer (CaP). These studies examined the antiproliferative effects of calcitriol in primary cell cultures derived from transgenic adenocarcinoma of mouse prostate (TRAMP) mice chronically treated with calcitriol (20 μg/kg) or vehicle 3×/week from 4 weeks-of-age until palpable tumors developed. This is a report on the response of two representative control (Vitamin D naïve, naïve) and calcitriol-treated (Vitamin D insensitive, VDI) cells to calcitriol. VDI cells were less sensitive to calcitriol based on less cell growth inhibition and less inhibition of DNA synthesis as measured by MTT and BrdU incorporation assays. Similarly, VDI cells were less sensitive to growth inhibition by the vitamin analog, 19-nor-1,25-dihydroxyvitamin D2 (paricalcitol). There was no change in apoptosis following treatment of naïve and VDI cells with calcitriol. Vitamin D receptor (VDR) expression was up-regulated by calcitriol in both naïve and VDI cells. In addition, calcitriol induced the Vitamin D metabolizing enzyme, 24-hydroxylase (cyp24) mRNA and enzyme activity similarly in naïve and VDI cells as measured by RT-PCR and HPLC, respectively. In summary, VDI cells are less responsive to the antiproliferative effects of calcitriol. Understanding Vitamin D insensitivity will further clinical development of Vitamin D compounds for prevention and treatment of CaP.  相似文献   

5.
6.
The rapid, non-genomic actions of 1,25-dihydroxyvitamin D(3) [1,25(OH)(2)D(3)] have been well described, however, the role of the nuclear vitamin D receptor (VDR) in this pathway remains unclear. To address this question, we used VDR(+/+) and VDR(-/-) osteoblasts isolated from wild-type and VDR null mice to study the increase in intracellular calcium ([Ca(2+)](i)) and activation of protein kinase C (PKC) induced by 1,25(OH)(2)D(3). Within 1 min of 1,25(OH)(2)D(3) (100 nM) treatment, an increase of 58 and 53 nM in [Ca(2+)](i) (n = 3) was detected in VDR(+/+) and VDR(-/-) cells, respectively. By 5 min, 1,25(OH)(2)D(3) caused a 2.1- and 1.9-fold increase (n = 6) in the phosphorylation of PKC substrate peptide acetylated-MBP(4-14) in VDR(+/+) and VDR(-/-) osteoblasts. The 1,25(OH)(2)D(3)-induced phosphorylation was abolished by GF109203X, a general PKC inhibitor, in both cell types, confirming that the secosteroid induced PKC activity. Moreover, 1,25(OH)(2)D(3) treatment resulted in the same degree of translocation of PKC-alpha and PKC-delta, but not of PKC-zeta, from cytosol to plasma membrane in both VDR(+/+) and VDR(-/-) cells. These experiments demonstrate that the 1,25(OH)(2)D(3)-induced rapid increases in [Ca(2+)](i) and PKC activity are neither mediated by, nor dependent upon, a functional nuclear VDR in mouse osteoblasts. Thus, VDR is not essential for these rapid actions of 1,25(OH)(2)D(3) in osteoblasts.  相似文献   

7.
The vitamin D receptor (VDR) is a member of the steroid receptor gene family. In this report, we examine the nature of specific VDR DNA binding utilizing the vitamin D-responsive element derived from the human osteocalcin promoter. Association of the VDR with the human osteocalcin 1,25-dihydroxyvitamin D3 (1,25(OH)2D3) responsive element (VDRE) in vitro was characterized on VDRE affinity columns by both weak and strong interactions. Weak interaction was a property of the VDR itself, monomeric in nature, and determined exclusively by the VDR's DNA-binding domain. Strong interaction, in contrast, was dependent upon an intact receptor molecule as well as a heterologous mammalian cell nuclear accessory factor (NAF). Heteromeric interaction between VDR and NAF was independent of the VDR DNA-binding domain, suggesting the presence of a functional dimerization domain separate from that for DNA binding. Direct association of NAF with immobilized VDR revealed that the interaction does not require the presence of DNA. Most importantly, while occupancy of the VDR by 1,25(OH)2D3 was not required for VDR interactions with either DNA or NAF, the presence of hormone increased the apparent relative affinity of the VDR for NAF approximately 10-fold. These studies suggest that high affinity association of the VDR with DNA requires both the DNA-binding domain as well as an additional independent structure located within the steroid-binding region. This protein subdomain interacts with NAF and is regulated by 1,25(OH)2D3.  相似文献   

8.
《The Journal of cell biology》1990,111(6):2385-2395
Prior studies have given no evidence for regulation of vitamin D receptor (VDR) compartmentalization or subcellular organization. Microwave fixation (9-15 s) and an indirect immunodetection system of avidin-biotin enhancement and phycoerythrin fluorophore resulted in sufficient spatial and temporal resolution to allow analysis of these processes. We studied cultured fibroblasts from normals or from patients with four different types of hereditary defect compromising VDR function (mutant cells). Compartmentalization of VDRs in the absence of 1,25-dihydroxyvitamin D3 (calcitriol) was regulated by serum or estrogen. VDRs were mainly cytoplasmic in cells cultured without serum and phenol red, but VDRs were mainly intranuclear after addition of serum or an estrogen to cells for at least 18 h (slow regulation). Calcitriol initiated a rapid and multistep process (rapid regulation) of reorganization in a portion of VDRs: clumping within 15-45 s, alignment of clumps along fibrils within 30-45 s, perinuclear accumulation of clumps within 45-90 s, and intranuclear accumulation of clumps within 1-3 min. We found similar rapid effects of calcitriol on VDRs in various other types of cultured cells. These sequential VDR pattern changes showed calcitriol dose dependency and calcitriol analogue specificity characteristic for the VDR. In mutant fibroblasts VDR pattern changes after calcitriol were absent or severely disturbed at selected steps. Treatment of normal cells with wheat germ agglutinin, which blocks protein transport through nuclear pores, also blocked calcitriol-dependent translocation of VDRs. We conclude that immunocytology after microwave fixation provides evidence for regulation of VDR organization and localization.  相似文献   

9.
UV exposure and serum levels of vitamin D have been linked in several studies with prostate cancer risk. At the cellular level, the principal action of vitamin D is mediated though vitamin D receptors (VDR). Since prostate cancer is a disease strongly associated with age, we examined the presence of VDR in normal prostate from donors of various ages to determine if the VDR expression pattern changed with age. We also compared the VDR expression in the peripheral and central zones of the prostate to determine if the expression pattern varied by location. Immunohistochemical studies were performed on paraffin-embedded tissue from cases selected by the following age decades; 10-19, 20-29, 30-39, 40-49, 50-59, and 60-69. Both the central and peripheral zones were examined for VDR expression. The intensity of VDR expression in prostate was compared with expression in different types of human tissues. Mean VDR expression was lowest in the 10-19 years of age group. The intensity of the nuclear VDR was higher though the fifth decade, and then declined in cases of ages 60-70. When multiple sections of the same donor prostate were compared, VDR expression was greater in the peripheral zone compared to the central zone.  相似文献   

10.
Others and we previously showed that the vitamin D receptor (VDR) is subject to degradation by the 26S proteasome and that treatment with 1,25‐dihydroxyvitamin D3 (1,25D3) inhibited this degradation. In the present study, we found that in osteoblasts, but not in intestinal epithelial cells, the VDR was susceptible to degradation by the 26S proteasome. The subcellular site for degradation of the VDR in osteoblasts is the cytoplasm and the site for ligand‐dependent protection of the VDR from the 26S proteasome is the chromatin. These direct relationships between nuclear localization and protection of the VDR from 26S proteasome degradation led us to hypothesize that the unoccupied cytoplasmic VDR is a substrate for polyubiquitination, which targets VDR for degradation by the 26S proteasome, and that nuclear localization has the ability to protect the VDR from polyubiquitination and degradation. To test these hypotheses, we used Cos‐1 cells transfected with human VDR and histidine‐tagged ubiquitin expression vectors. We found that unoccupied VDR was polyubiquitinated and that 1,25D3 inhibited this modification. Mutations in the nuclear localization signal of VDR (R49W/R50G and K53Q/R54G/K55E) or in the dimerization interface of VDR with retinoid X receptor (M383G/Q385A) abolished the ability of 1,25D3 to protect the VDR from polyubiquitination, although these mutations had no effect on the ligand‐binding activity of VDR. Therefore, we concluded that in some cellular environments unoccupied cytoplasmic VDR is susceptible to polyubiquitination and proteasome degradation and that ligand‐dependent heterodimerization and nuclear localization protect the VDR from these modifications. J. Cell. Biochem. 110: 926–934, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

11.
12.
We have previously purified a cytosolic vitamin D metabolite binding protein (cDBP) from rat enterocytes, which has characteristics distinct from other vitamin D binding proteins. In these studies, we demonstrate that cDBP in a semi-purified fraction from human intestinal cells (Caco-2 cells) binds 25-hydroxyvitamin D (25OHD) with at least a 1000-fold greater affinity than 1, 25-dihydroxyvitamin D (1,25(OH)(2)D) or 24,25-dihydroxyvitamin D. Treatment of cells with 1,25(OH)(2)D reduced 25OHD binding to approximately one third that of the untreated cells (0.42 CPM/mg total protein vs 1.34 CPM/mg total protein, respectively). Finally, the cDBP is not immunoreactive to antibodies prepared against the C-terminus of the nuclear vitamin D receptor (VDR). In summary, cDBP bound 25OHD with greater affinity than either 1,25(OH)(2)D or 24,25 dihydroxyvitamin D, the cytosolic binding activity was down-regulated by 1,25(OH)(2)D and cBDP is distinct from the nuclear VDR.  相似文献   

13.
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   

14.
15.
We undertook an investigation of an outbreak of rachitic bone disease in the Emperor Tamarin New World primate colony at the Los Angeles Zoo in the mid-1980s. The disease phenotype resembled that observed in humans with an inactivating mutation of the vitamin D receptor (VDR), hypocalcemia, high 1,25-dihydroxyvitamin D (1,25-(OH)(2)D) levels, and rickets in rapidly growing adolescent primates. In contrast to the human disease, the New World primate VDR was functionally normal in all respects. The proximate cause of vitamin D hormone resistance in New World primates was determined to be the constitutive overexpression of a heterogeneous nuclear ribonucleoprotein in the A family which we coined the vitamin D response element binding protein (VDRE-BP). VDRE-BP competed in trans with the VDR-retinoid X receptor (RXR) for binding to the vitamin D response element. VDRE-BP-legislated resistance to 1,25-(OH)(2)D was antagonized (i.e., compensated) by another set of constitutively overexpressed proteins, the hsp-70-related intracellular vitamin D binding proteins (IDBPs). IDBPs, present but expressed at much lower levels in Old World primates including man, exhibited a high capacity for 25-hydroxylated vitamin D metabolites and functioned to traffic vitamin Ds to specific intracellular destinations to promote their action and metabolism.  相似文献   

16.
17.
We recently reported (Larsson and Nemere [2003]: Endocrinology 144:1726) the effects of growth and maturation on 1,25(OH)2D3-membrane initiated effects in the intestine of male chickens. Here we extend our observations to studies on females with two stages of high calcium demand: growth (7-14 weeks) and egg laying (28-58 weeks). The rapid stimulatory effect of 130 pM 1,25(OH)2D3 on calcium transport was assessed as a physiological response in perfused duodena of 7-, 14-, 28-, and 58-week-old chickens, and determined to be 308%, 184%, 170%, and 153%, respectively, of corresponding controls after 40 min. Saturation analyses of [3H]1,25(OH)2D3 binding to nuclear vitamin D receptor (VDR) indicated an absence of cooperative binding, no changes in dissociation constant (Kd) with age, and an increase in maximum binding capacity (Bmax) between 7-week birds and older age groups. Analyses of saturable binding of [3H]1,25(OH)2D3 to the membrane associated rapid response steroid binding protein (1,25D3-MARRS bp) in basal lateral membranes (BLM), indicated cooperative binding, and an increase in both Bmax and Kd with age. No changes in the age-related expression of 1,25D3-MARRS bp were found, as judged by Western analyses, suggesting that a shift in ligand binding to lower affinity membrane components accounted for the increase in calculated Bmax. Basal levels of protein kinase C (PKC) activity decreased with age, as did hormone enhancement of activity. Basal levels of protein kinase A (PKA) activity remained constant with age, while the magnitude of hormone stimulation increased. Comparison of dose-response curves for ion transport and kinase activities in 7-week chicks suggested that PKC mediates phosphate transport while PKA mediates calcium transport. Thus, the age-related loss of calcium transport is most likely related to loss of PKC-mediated phosphate transport.  相似文献   

18.
19.
20.
The vitamin D receptor (VDR) is a member of the steroid/retinoid receptor superfamily of nuclear receptors and has potential tumor-suppressive functions in prostate and other cancer types. Vitamin D3 (VD3) exerts its biological actions by binding within cells to VDR. The VDR then interacts with specific regions of the DNA in cells, and triggers changes in the activity of genes involved in cell division, cell survival, and cellular function. Using human primary cultures and the prostate cancer (PCa) cell line, ALVA-31, we examined the effects of VD3 under different culture conditions. Complete G0/G1 arrest of ALVA-31 cells and approximately 50% inhibition of tumor stromal cell growth was observed. To determine changes in gene expression patterns related to VD3 activity, microarray analysis was performed. More than approximately 20,000 genes were evaluated for twofold relative increases and decreases in expression levels. A number of the gene targets that were up- and down-regulated are related to potential mechanisms of prostatic growth regulation. These include estrogen receptor (ER), heat shock proteins: 70 and 90, Apaf1, Her-2/neu, and paxillin. Utilizing antibodies generated against these targets, we were able to confirm the changes at the protein level. These newly reported gene expression patterns provide novel information not only potential markers, but also on the genes involved in VD3 induced apoptosis in PCa.  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