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.
The C(23) epimers of 1alpha,23,25(OH)3-24-oxovitamin D3, a major natural metabolite of the secosteroid hormone, 1alpha,25(OH)2D3, were chemically synthesized for the first time. The metabolite was synthesized by palladium coupling of the appropriate CD ring analog with an A ring enyne. Various approaches from quinic acid to the A ring precursors were explored, and a new route to the A ring enyne from quinic acid was developed. The C(23) stereochemistry of the natural 1alpha,23,25(OH)3-24-oxovitamin D3 produced in neonatal human keratinocytes was determined to be S on the basis of the 1H NMR and the HPLC data. The biological activity of 1alpha,23(S), 25(OH)3-24-oxovitamin D3 in primary cultures of bovine parathyroid cells was determined by comparing the potency of this metabolite to that of 1alpha,25(OH)2D3 in suppression of parathyroid hormone (PTH) secretion. The results indicate that 1alpha,23(S), 25(OH)3-24-oxovitamin D3 potently suppressed PTH secretion even at concentrations as low as 10(-)12 M and is equipotent with 1alpha, 25(OH)2D3. The high activity of 1alpha,23(S),25(OH)3-24-oxovitamin D3 cannot be explained on the basis of its affinity for the vitamin D receptor as this metabolite was found to be 10 times less effective than radioinert 1alpha,25(OH)2D3 in blocking the uptake and receptor binding of [3H]-1alpha,25(OH)2D3 in intact parathyroid cells. Further studies are required to explain the molecular basis for the activity of 1alpha,23(S),25(OH)3-24-oxovitamin D3 in its ability to suppress PTH secretion. In summary, our present study indicates that the C(23) stereochemistry of the natural 1alpha,23, 25(OH)3-24-oxovitamin D3 is S and this metabolite is equipotent to 1alpha,25(OH)2D3 in suppressing PTH secretion.  相似文献   

2.
Previous biological studies have implicated two vitamin D metabolites, 1 alpha,25(OH)2-vitamin D3[1 alpha,25(OH)2-D3] and 24R,25(OH)2-vitamin D3 [24R,25(OH)2D3] in the process of skeletal fracture-healing. While a nuclear receptor for 1 alpha,25(OH)2D3 is known to be present in osteoblast and absent in osteoclast cell lines, no systematic study has been carried out on the callus tissue which is formed during fracture-healing. The present report shows that a binding protein/receptor for 1 alpha,25(OH)2D3 resides both in a postnuclear membrane fraction and in a high speed cytosol fraction of the callus tissue obtained 10 days after imposition of a tibial fracture. The dissociation constant, KD, for 1 alpha,25(OH)2D3 was 0.83 +/- 0.34 M and 0.66 +/- 0.38 nM respectively, for the membrane and cytosol fractions. Results from a panel of steroid competition assays indicate that both receptor/binding proteins greatly prefer 1 alpha-hydroxylated ligands as compared to 1 alpha-deoxy or 24-hydroxylated ligands. The presence of 1 alpha,25(OH)2D3 receptors in the fracture-healing callus is consistent with the known biological effects of the metabolite on the fracture-healing process.  相似文献   

3.
The hormonally active form of vitamin D, 1 alpha,25-dihydroxyvitamin D3 [1 alpha,25(OH)2D3] stimulates biological responses related to calcium homeostasis, cell differentiation, and immunomodulation in many target cells, including leukemic cells. Most of these responses are dependent upon 1 alpha,25(OH)2D3 interaction with a nuclear receptor protein. Structural analogues of 1 alpha,25(OH)2D3 might allow for separation of biological function, avoiding adverse calcemic effects. This report quantitates intestinal calcium absorption, bone calcium resorption, induction of intestinal and renal calcium-binding protein (CaBP), and occupancy of the intestinal and renal nuclear 1 alpha,25(OH)2D3 receptor in vitamin D-deficient chicks after a single dose of 1 alpha,25(OH)2D3, 1 alpha,25-dihydroxyvitamin-16-ene-23-yne-D3 (analogue V), or 22-[m-(dimethylhydroxymethyl)phenyl]-23,24,25,26,27- pentanor-1 alpha-hydroxy-vitamin D3 (analogue EV). The interaction of these compounds with chick intestinal nuclear 1 alpha,25(OH)2D3 receptor and chick plasma vitamin D-binding protein was determined in vitro; analogues V and EV bound 68% and 62% [1 alpha,25(OH)2D3 receptor] and 8% and 13% (vitamin D-binding protein), respectively, as well as 1 alpha,25(OH)2D3 (100%). 1 alpha,25(OH)2D3 doses (0.075-1.2 nmol) generated responses in intestinal calcium absorption, bone calcium resorption, intestinal CaBP, and renal CaBP. When analogue V (1.2-300 nmol) was administered, increases in bone calcium resorption and renal CaBP were noted. However, a significant response in intestinal calcium absorption and intestinal CaBP appeared only after a 300-nmol dose. Unoccupied nuclear 1 alpha,25(OH)2D3 receptor in the intestine and kidney was determined in vivo after doses of 1 alpha,25(OH)2D3, analogue V, or analogue EV. Doses (0.25-6.0 nmol) of 1 alpha,25(OH)2D3 and analogue EV reduced unoccupied receptor to 24% and 59% (intestine) and to 13% and 41% (kidney), respectively. Analogue V (6.0-600 nmol) decreased unoccupied receptor in the kidney. In the intestine analogue V (300-600 nmol) reduced unoccupied receptor only to 75%. These results confirm that some vitamin D analogues can generate selective biological responses and different levels of target organ receptor occupancy.  相似文献   

4.
Vitamin D3 and its hydroxylated metabolites are normally in thermal equilibrium with their previtamin D isomers. To evaluate the biologic activity of 1 alpha, 25-dihydroxyprevitamin D3, we synthesized 19-nor analogs of 1 alpha, 25-dihydroxy(pre)vitamin D3 because the absence of a C19 methylene group prevents the isomerization of these analogs. The affinity of 1 alpha, 25-(OH)2D3-19-nor-D3 for the intestinal vitamin D receptor and plasma vitamin D binding protein was mildly decreased [30 and 20% of the affinity of 1 alpha, 25-(OH)2D3, respectively], but the affinity of 1 alpha, 25-(OH)2-19-nor-previtamin D3 was only 1 and 6% of that of 1 alpha, 25-(OH)2D3 for the receptor and DBP, respectively. The in vitro effects on human promyeloid leukemia (HL-60 cell) differentiation and osteocalcin secretion by human osteosarcoma (MG-63) cells by 1 alpha, 25-(OH)2-19-nor-D3 were nearly identical to those of 1 alpha-25-(OH)2D3, whereas 19-nor-previtamin D3 showed poor activity (2%). The in vivo calcemic effects of both analogs, studied in vitamin D-deficient chicks treated for 10 consecutive days with the analogs, showed no activity of the previtamin D3 analog and reduced calcemic effects (< or = 10%) of 1 alpha, 25-(OH)2-19-nor-D3. We conclude that the previtamin D form of 1 alpha, 25-(OH)2D3 has lost most of its biologic activity in vitro and in vivo.  相似文献   

5.
The steroid hormone 1 alpha,25-dihydroxyvitamin D3 [1 alpha,25(OH)2D3] stimulates biological responses via both genomic mechanisms and nongenomic mechanisms (opening of voltage-gated Ca2+ channels). We report here that the two closed B-ring steroid analogs of 1 alpha,25(OH)2D3, 1 alpha,25(OH)2-7-dehydrocholesterol and 1 alpha,25(OH)2-lumisterol3, are able to generate the nongenomic response, transcaltachia, without the ability to compete with the natural metabolite for binding to its nuclear receptor. We propose that the nongenomic membrane associated receptor can accept the ligand in its closed "6-s-cis" conformation whereas the nuclear receptor prefers the extended "6-s-trans" conformer.  相似文献   

6.
Previous in vivo studies have shown that growth hormone (GH) affects vitamin D and mineral metabolism. Insulin-like growth factor-I (IGF-I) was recently reported to be a regulator of renal 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3) production, suggesting that it mediates the effects of GH on vitamin D metabolism. However, there is no direct evidence to support this. The present study was designed to investigate the in vitro effects of GH and IGF-I on the renal production of 1,25-(OH)2D3 and 24,25-dihydroxyvitamin D3 (24,25-(OH)2D3) in a pig kidney cell line, LLC-PK1. Confluent cells were preincubated in serum-free medium with hormone (GH or IGF-I) or vehicle, and then incubated with 25-[3H]OHD3. The levels of 1,25-[3H](OH)2D3 and 24,25-[3H](OH)2D3 produced were determined after lipid extraction and HPLC purification. Production of 1,25-(OH)2D3 and 24,25-(OH)2D3 was increased after both IGF-I and GH preincubation in a dose-dependent manner. Significant increases were found after preincubation with 13 nmol/l IGF-I (1,25-(OH)2D3, 1.8-fold: 24,25-(OH)2D3, 1.5-fold)or 0.9 or 9 nmol/lGH (1,25-(OH)2D3, 1.3-fold and 1.5-fold; 24.25-(OH)2D3, 1.4-fold and 1.5-fold respectively). Furthermore, the effect of 9 nmol/l GH on 1.25-(OH)2D3 and 24,25-(OH)2D3 production was blocked in the presence of IGF-I receptor monoclonal antibody. These results confirm that IGF-I acts on renal tubules, resulting in induction of 1,25-(OH)2D3 and 24,25-(OH)2D3 production, and the findings suggest that GH stimulates 1.25-(OH)2D3 and 24,2 5-(OH)2D3 production by increasing local IGF-I production in the kidney.  相似文献   

7.
Matrix vesicles are extracellular organelles produced by cells that mineralize their matrix. They contain enzymes that are associated with calcification and are regulated by vitamin D metabolites in a cell maturation-dependent manner. Matrix vesicles also contain metalloproteinases that degrade proteoglycans, macromolecules known to inhibit calcification in vitro, as well as plasminogen activator, a proteinase postulated to play a role in activation of latent TGF-beta. In the present study, we examined whether matrix vesicle metalloproteinase and plasminogen activator are regulated by 1, 25(OH)2D3 and 24,25(OH)2D3. Matrix vesicles and plasma membranes were isolated from fourth passage cultures of resting zone chondrocytes that had been incubated with 10(-10)-10(-7) M24, 25(OH)2D3 or growth zone chondrocytes incubated with 10(-11)-10(-8) M 1,25(OH)2D3, and their alkaline phosphatase, active and total neutral metalloproteinase, and plasminogen activator activities determined. 24,25(OH)2D3 increased alkaline phosphatase by 35-60%, decreased active and total metalloproteinase by 75%, and increased plasminogen activator by fivefold in matrix vesicles from resting zone chondrocyte cultures. No effect of vitamin D treatment was observed in plasma membranes isolated from these cultures. In contrast, 1,25(OH)2D3 increased alkaline phosphatase by 35-60%, but increased active and total metalloproteinase three- to fivefold and decreased plasminogen activator by as much as 75% in matrix vesicles isolated from growth zone chondrocyte cultures. Vitamin D treatment had no effect on plasma membrane alkaline phosphatase or metalloproteinase, but decreased plasminogen activator activity. The results demonstrate that neutral metalloproteinase and plasminogen activator activity in matrix vesicles are regulated by vitamin D metabolites in a cell maturation-specific manner. In addition, they support the hypothesis that 1,25(OH)2D3 regulation of matrix vesicle function facilitates calcification by increasing alkaline phosphatase and phospholipase A2 specific activities as well as metalloproteinases which degrade proteoglycans.  相似文献   

8.
It has recently been shown that in the rat, dihydrotachysterol (DHT) is extensively metabolized in the side-chain in vivo along pathways similar to those of vitamin D. In addition 25-hydroxy-DHT2 [25OHDHT2] is hydroxylated at C1, producing both 1 alpha- and 1 beta- hydroxy compounds. An in vivo study in 1988 demonstrated that in normal adult subjects receiving oral DHT2, plasma 1 alpha,25-dihydroxyvitamin D [1,25-(OH)2D] concentrations fell, but with unchanged plasma PTH levels. Down-regulation of 1,25-(OH)2D3 production by 25-(OH)DHT2 or some other unknown metabolite was also suggested as an explanation for these observations. To investigate whether either of the newly characterized 1 alpha,25- or 1 beta,25-(OH)2DHT2 was formed in vivo in normal man, DHT2 (approximately 1 mg/day, orally) was administered to healthy volunteers (three males and one female). Plasma was analyzed by high performance liquid chromatography and gas chromatography-mass spectrometry, demonstrating the formation of both 1 alpha,25- and 1 beta,25-(OH)2DHT2 in vivo in normal human subjects. Plasma levels of 1,25-(OH)2D3, PTH, ionized and total calcium, inorganic phosphate, and alkaline phosphatase were monitored. The plasma concentrations of DHT2, 25OHDHT2, and 1 alpha,25- and 1 beta,25-(OH)2DHT2 were measured by gas chromatography-mass spectrometry. In all volunteers, plasma ionized calcium increased slightly during DHT2 administration; 1,25-(OH)2D3 and PTH concentrations fell. Plasma levels of DHT2 and its metabolites rose over the same period. The average fall in the level of plasma 1,25-(OH)2D (60-70 pmol/L) was mirrored by a rise in the concentration of 1 alpha,25-(OH)2DHT2 (550 pmol/L). This ratio is appropriate, because it has previously been shown that in a reconstituted COS cell, 1 alpha,25-(OH)2DHT3 has roughly one tenth the potency of 1,25-(OH)2D3. At maximum concentration, the ratios of DHT2/25OHDHT2/1 beta,25-(OH)2DHT2/1 alpha,25-(OH)2DHT2 were approximately 10:1:2:0.1. The concentration of 1 beta,25-(OH)2DHT2 was greater than that of 25OHDHT2, and the ratio of 1 alpha,25- to 1 beta,25-(OH)2DHT2 (1:20) was substantially lower than that in rat plasma (3:10). The data presented here suggest that the active DHT2 metabolite in man is 1 alpha,25-(OH)2DHT2 and that the fall in plasma 1,25-(OH)2D seen during DHT therapy may be partly the result of suppressed PTH secretion.  相似文献   

9.
In mouse bone marrow primary cultures, the formation of osteoclast-like, i.e. tartrate-resistant acid phosphatase (TRAP)- and calcitonin receptor-positive multinucleated cells (MNC), when induced by 1alpha,25-dihydroxyvitamin D3 (1alpha,25(OH)2D3), can be suppressed by 17beta-estradiol (17beta-E2), whereas 17alpha-E2 is without any effect. 17beta-E2, above 10(-11) M, significantly reduced 1alpha,25(OH)2D3-mediated TRAP+ MNC formation in cultured bone marrow cells from both female and male mice. The estrogen at 10(-8) M suppressed the peak response to the vitamin D sterol by 50%. 17beta-E2 significantly suppressed basal and 1alpha,25(OH)2D3-stimulated cellular production of interleukin (IL)-6. IL-6 alone, although bone marrow cells in hormone-free culture produced appreciable amounts of the cytokine, did not induce any TRAP+ MNC. Therefore, the changes in IL-6 production induced by the hormones could not be the sole determinant for the extent of TRAP+ MNC formation. However, the stimulatory effect of 1alpha,25(OH)2D3 on osteoclastogenesis nevertheless can be significantly reduced by a neutralizing monoclonal anti-IL-6 antibody. In the presence of 10(-8) M 17beta-E2, the anti-IL-6 monoclonal antibody does not achieve any further suppression of 1alpha,25(OH)2D3-related osteoclast-like cell formation. Our data suggest that induction of osteoclastogenesis by 1alpha,25(OH)2D3 is partially dependent on IL-6 signaling and can be modulated by 17beta-E2 through interference with IL-6 receptor activation, in addition to inhibition of IL-6 production by marrow stromal cells.  相似文献   

10.
1alpha,25-Dihydroxyvitamin D3 [1alpha,25(OH)2D3] regulates cellular growth and differentiation. We show that in keratinocytes, 1alpha, 25(OH)2D3 reduces concentrations of the messenger RNA of IEX-1, the product of which blocks Fas- or tumor necrosis factor type alpha-induced apoptosis in various cells. In sub-confluent keratinocyte cultures, the addition of 1alpha,25(OH)2D3, in amounts that induce growth arrest, reduces IEX-1 mRNA concentrations. In confluent cells, 1alpha,25(OH)2D3 initially reduces and then increases IEX-1 mRNA concentrations. IEX-1 protein is localized in the nucleus and perinuclear region of keratinocytes. In sub-confluent cells, 1alpha,25(OH)2D3 translocates IEX-1 protein from the nucleus to the perinuclear region and cytoplasm. Since IEX-1 has recently been shown to regulate cell survival and number, we suggest that IEX-1 may play a role in keratinocyte growth and differentiation and that 1alpha,25(OH)2D3 may reduce keratinocyte growth via a reduction in IEX-1 mRNA and a change in the intracellular distribution of IEX-1 protein.  相似文献   

11.
12.
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 (1 alpha-(hydroxymethyl)-3 beta-hydroxy-20-epi-22-oxa-26,27-dihomo vitamin D3 (analogue MCW-YA = 3a) and 1 beta-(hydroxymethyl)-3 alpha-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.  相似文献   

13.
Although prostatic cancer is often viewed as an androgen-dependent malignancy, a number of other hormones including 1alpha, 25-dihydroxyvitamin D3 [1alpha,25(OH)2D3] are now recognized to modulate its growth and differentiated phenotype. Seven different continuous human prostatic carcinoma cell lines were examined for the presence of biologically active receptors for 1alpha,25(OH)2D3. All seven lines were found to contain mRNA for the vitamin D receptor using an RNase protection assay. Six of the seven cell lines were found to have high-affinity saturable binding sites for 1alpha,25(OH)2D3. The seventh line was found to contain vitamin D receptors by sucrose gradient analysis. All seven lines were found to express 24-hydroxylase activity by a HPLC assay that measures the conversion of 25-hydroxyvitamin D3 to 24,25-dihydroxyvitamin D3. 24-Hydroxylase activity was up-regulated in all seven cell lines by preincubation with 1alpha,25(OH)2D3. In the presence of fetal bovine serum, the growth of four of the seven cell lines was inhibited. In the majority of cell lines growth inhibition was related not only to the number of receptors per cell, but also in inverse proportion to the 24-hydroxylase activity of each cell line. The ubiquitous presence of vitamin D receptor and 24-hydroxylase activity in human prostatic carcinoma cells suggests new alternatives for the pharmacological treatment of advanced prostatic cancer and implies that chemoprevention strategies could also make use of this endocrine axis.  相似文献   

14.
Cytosol binding protein for 1alpha,25-(OH)2D3 and 25-OHD3 was demonstrated in rat intestinal mucosa using sucrose density gradient ultracentrifugation analysis, Sephadex G-200 column chromatography and polyacrylamide disc gel electrophoresis. The binding protein has a sedimentation constant of 5-6S, a molecular weight of 100,000-120,000 daltons and a mobility in electrophoresis of Rf 0.42. Further analysis of binding behavior by DEAE-cellulose filter assays revealed that the bindig protein had a higher affinity for 25-OHD3 than for 1alpha,25-(OH)2D3, and that 1alpha,25-(OH)2D3 competed with 25-OHD3 for the same binding site on the protein. The apparent dissociation constant for 1alpha,25-(OH)2D3 and 25-OHD3 was 9.2 X 10(-9) M and 1.5 X 10(-9) M, respectively. The binding capacity was 3.1 pmoles/mg protein for both 1alpha,25-(OH)2D3 and 25-OHD3. The order of binding affinity was 25-OHD3 greater than 1alpha,25-(OH)2D3 greater than D3 = 1alpha-OHD3.  相似文献   

15.
26,26,26,27,27,27-Hexafluo-1,25(OH)2 vitamin D3, the hexafluorinated analog of 1,25(OH)2 vitamin D3, has been reported to be several times more potent than the parent compound regarding some vitamin D actions. The reason for enhanced biologic activity in the kidneys and small intestine appears to be related to F6-1,25(OH)2 vitamin D3 metabolism to ST-232, 26,26,26,27,27,27-hexafluoro-1 alpha, 23S,25-trihydroxyvitamin D3, a bioactive 23S-hydroxylated form that is resistant to further metabolism. Since F6-1,25(OH)2 vitamin D3 is considered to prevent osteoporotic decrease in bone mass by suppressing bone turnover, we here compared the distribution and metabolism of [1 beta-3H]F6-1,25(OH)2 vitamin D3 and [1 beta-3H]1,25(OH)2 vitamin D3 in bones of rats by autoradiography and radio-HPLC. In the dosed groups, radioactivity was detected locally in the metaphysis, the modeling site in bones. As compared with the [1 beta-3H]1,25(OH)2 vitamin D3 case, [1 beta-3H]F6-1,25(OH)2 vitamin D3 was significantly retained in this site, and moreover, it mainly persisted as unchanged compound and ST-232. These findings indicate that the reason for the higher potency of F6-1,25(OH)2 vitamin D3 than 1,25(OH)2 vitamin D3 in bones are linked with increased distribution and reduced metabolism.  相似文献   

16.
Epidemiological and laboratory data support a role for vitamin D in the growth and differentiation of human prostatic cells. These findings prompted us to ask whether prostatic cells could convert 25-hydroxyvitamin D3 (25-OH-D3), the major circulating metabolite of vitamin D3, to 1,25-dihydroxyvitamin D3 [1,25(OH)2D3], the hormonally active metabolite, in a manner similar to cultured human keratinocytes. Therefore, we investigated three well-characterized human prostate cancer cell lines, LNCaP, DU 145, and PC-3; two primary cultures of cells derived from noncancerous human prostates (one normal and one benign prostatic hyperplasia); and primary cultures of normal human keratinocytes for their ability to synthesize 1,25(OH)2D3. Assays were performed in the presence of 25-OH-D3 as the enzyme substrate and 1,2-dianilinoethane, an antioxidant and free radical scavenger, and in the presence and absence of clotrimazole, a cytochrome P450 inhibitor. DU 145 and PC-3 cells produced 0.31 +/- 0.06 and 0.07 +/- 0.01 pmol of 1,25(OH)2D3/mg protein/h, respectively. No measurable 1,25(OH)2D3 was detected in LNCaP cells. The normal and benign prostatic hyperplasia primary cultures and keratinocyte cultures produced 3.08 +/- 1.56, 1.05 +/- 0.31, and 2.1 +/- 0.1 pmol of 1,25(OH)2D3/mg protein/h, respectively, using a calf thymus receptor binding assay to measure 1,25(OH)2D3 in the presence of 1,2-dianilinoethane. The identity of the analyte as 1,25(OH)2D3 was supported by high performance liquid chromatography using [3H]25-OH-D3 as the enzyme substrate and a solvent system that is specific for 1,25(OH)2D3. The production of 1,25(OH)2D3 in the prostate cancer cell lines and in the primary cultures was completely inhibited in the presence of clotrimazole. This report demonstrates that two of three human prostate cancer cell lines, as well as primary cultures of noncancerous prostatic cells, possess 1alpha-hydroxylase activity and can synthesize 1,25(OH)2D3 from 25-OH-D3. Together with recent data indicating that 1,25(OH)2D3 inhibits the invasiveness of human prostate cancer cells (G. G. Schwartz et al., Cancer Epidemiol. Biomark. Prev., 6: 727-732, 1997), these data suggest a potential role for 25-OH-D3 in the chemoprevention of invasive prostate cancer.  相似文献   

17.
The hormone 1alpha,25-dihydroxyvitamin D3 [1alpha,25(OH)2D3] generates biological responses via both genomic and rapid, nongenomic mechanisms. The genomic responses utilize signal transduction pathways linked to a nuclear receptor (VDRnuc) for 1alpha,25(OH)2D3, while the rapid responses are believed to utilize other signal transduction pathways that may be linked to a putative membrane receptor for 1alpha,25(OH)2D3. The natural seco steroid is capable of facile rotation about its 6,7 single carbon bond, which permits generation of a continuum of potential ligand shapes extending from the 6-s-cis (steroid like) to the 6-s-trans (extended). To identify the shape of conformer(s) that can serve as agonists for the genomic and rapid biological responses, we measured multiple known agonist activities of two families of chemically synthesized analogs that were either locked in the 6-s-cis (6C) or 6-s-trans (6T) conformation. We found that 6T locked analogs were inactive or significantly less active than 1alpha,25(OH)2D3 in both rapid responses (transcaltachia in perfused chick intestine, 45Ca2+ influx in ROS 17/2.8 cells) and genomic (osteocalcin induction in MG-63 cells, differentiation of HL-60 cells, growth arrest of MCF-7 cells, promoter transfection in COS-7 cells) assays. In genomic assays, 6C locked analogs bound poorly to the VDRnuc and were significantly less effective than 1alpha,25(OH)2D3 in the same series of assays designed to measure genomic responses. In contrast, the 6C locked analogs were potent agonists of both rapid response pathways and had activities equivalent to the conformationally flexibile 1alpha,25(OH)2D3; this represents the first demonstration that 6-s-cis locked analogs can function as agonists for vitamin D responses.  相似文献   

18.
Relaxin stimulates in vitro DNA synthesis and cell proliferation of porcine granulosa cells (GC) and theca cells. The objective of the study reported here was to determine whether components of the ovarian insulin-like growth factor (IGF) system mediate relaxin's growth-promoting effects on porcine GC in vitro. In small follicle GC, relaxin (1-100 ng/ml) significantly (p < 0.05) increased IGF-I secretion to 25-34% above control. Hormonal responsiveness of GC was shown by incubation with FSH (200 ng/ml), which resulted in 125% stimulation of IGF-I secretion relative to that in cells incubated alone. When IGF-I activity in the GC cultures was neutralized with a specific IGF-I antibody, relaxin (10 and 100 ng/ml)-induced [3H]thymidine incorporation was inhibited (p < 0.05). Coincubation with IGF-I antibody also suppressed basal and IGF-I (10 ng/ml)-induced [3H]thymidine incorporation into GC DNA, but had no effect on insulin (1 microgram/ml)-induced DNA synthesis, demonstrating the specificity and lack of toxicity of the IGF-I antibody. Ligand blot analysis showed no change in secretion of GC IGF binding protein (IGFBP) in response to relaxin (1, 10, and 100 ng/ml). In contrast, IGF-I (10 ng/ml) increased secretion of IGFBP-3 and -5, whereas FSH (200 ng/ml) decreased IGFBP-3 secretion and increased IGFBP-4 secretion (p < 0.05). In IGF-I receptor competition studies, IGF-I, but not relaxin, displaced [125I]IGF-I from the GC IGF-I receptor. These studies provide direct evidence for an interaction of relaxin and the ovarian IGF system. They are the first to show 1) a stimulatory effect of relaxin on IGF-I secretion; 2) the necessity of IGF-I activity for relaxin-induced GC DNA synthesis; and 3) the absence of an effect of relaxin on GC IGFBPs or IGF-I receptor. These findings support a paracrine role for relaxin in the porcine follicle and show that relaxin acts indirectly to promote follicle growth by stimulating GC IGF-I secretion.  相似文献   

19.
Effects of dexamethasone and 1,25-dihydroxyvitamin D3 [1,25(OH)2D3] were studied in cultures of adult human marrow stromal cells. In primary culture, dexamethasone (10(-8) M) increased the number of fibroblast colonies formed but decreased their average size. The number of colonies expressing alkaline phosphatase activity was increased, consistent with the enhancement of osteogenic differentiation by this glucocorticoid. In secondary culture, osteogenic differentiation was assessed by measurement of the steady-state levels of particular mRNAs that are characteristic of cells of the osteoblast lineage. The mRNAs for alpha 1(I)-procollagen, alkaline phosphatase, osteopontin and bone sialoprotein were expressed under all culture conditions used. In contrast, osteocalcin mRNA expression was detectable only in cultures treated with 1,25(OH)2D3 (10(-8) M). Addition of 1,25(OH)2D3 to control increased the expression of the mRNAs for alkaline phosphatase and osteopontin but had no significant effect on bone sialoprotein expression. The highest levels of expression of the mRNAs for alkaline phosphatase, bone sialoprotein and osteocalcin were observed in dexamethasone-treated cultures to which 1,25(OH)2D3 had been added. These results demonstrate that, as earlier found in other species, dexamethasone and 1,25(OH)2D3 promote the osteogenic differentiation of human marrow stromal cells as measured by expression of these osteogenic markers.  相似文献   

20.
After intravenous administration of the vitamin D3 analog, 22-oxacalcitriol (OCT), to normal rats plasma metabolites were investigated by HPLC, GC-MS and LC-MS. Five side-chain oxidation metabolites, 24R(OH)OCT, 24S(OH)OCT, (25R)-26(OH)OCT, (25S)-26(OH)OCT and 24oxoOCT, were identified by comparison with the corresponding synthetic compounds. These side-chain oxidation metabolites were similar to those of calcitriol [1alpha,25(OH)2 vitamin D3] described previously. Besides these five metabolites, two unique side-chain cleavage metabolites, 20S(OH)-hexanor-OCT and 17,20S(OH)2-hexanor-OCT, were identified as main metabolites in plasma by GC-MS and LC-MS using a specific chemical reaction. Our studies suggest that OCT is extensively metabolized and circulates in blood as a number of metabolites as well as unchanged OCT. This metabolism includes both unique pathways of C23-O22 cleavage and 17-hydroxylation, in addition to the side-chain oxidation metabolites similar to those of 1alpha,25-(OH)2D3.  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