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.
A new metabolite of Vitamin D3 (25-OHD3-26,23-lactone) has been found in the plasma of Vitamin D3-toxic pigs and cows. This metabolite is at least 5 times more potent than 25-OHD3 in the displacement of [3H]-25-OHD3 from rat plasma protein binding sites under short-term incubation. This metabolite co-migrates with 24,25-(OH)2D3 on Sephadex LH-20 columns developed in chloroform:hexane 65:35 and with 25,26-(OH)2D3 on Sephadex LH-20 columns developed in hexane:chloroform:methanol 9:1:1. The presence of 25-OHD3-26,23-lactone represents a possible contaiminant in the assay of 24,25-(OH)2D3 or 25,26-(OH)2D3 if only Sephadex LH-20 is used for pre-assay purification. 25-OHD3-26,23-lactone is, however, resolved from 24,25-(OH)2D3 by high pressure liquid chromatography (HPLC) using Zorbax Sil silicic acid columns developed in either isopropanol:hexane 8:92 or isopropanol:methylene chloride 2.5:96.5. We assayed for the presence of this new metabolite of Vitamin D3 and found it to be present in normal pig plasma and undetectable in normal cow plasma. Concentrations were elevated to 10–20 ng/ml following massive injection of Vitamin D3 to both species.  相似文献   

2.
The four stereoisomers of 25-hydroxyvitamin D3-26,23 lactone (25-OHD3-26,23 lactone) were tested against in vivo 25-OHD3-26,23 lactone to determine their relative competition in the radioligand binding assays for 25-OHD3 and 1,25-(OH)2D3. The 25R-OHD3-26,23S lactone and in vivo 25-OHD3-26,23 lactone behaved identically in the radioligand binding assay for 25-OHD3 and were ~5-fold more potent than 25-OHD3 at displacing 25-OH[3H]D3. The 25S-OHD3-26,23S lactone was the poorest competitor in this assay, requiring a 10-fold excess relative to 25-OHD3 to displace 50% of the 25-OH[3H]D3. The order of competition in the 25-OHD3 radioligand binding assay was 25R-OHD3-26,23S lactone = in vivo 25-OHD3-26,23 lactone ? 25S-OHD3-26,23R lactone > 25-OHD3 ? 25R-OHD3-26,23R lactone > 25S-OHD3-26,23S lactone. The order of competition in the 1,25-(OH)2D3 cytosol receptor assay was essentially reversed from the competition in the 25-OHD3 assay and was 25S-OHD3-26,23S lactone > 25-OHD3 ? 25S-OHD3-26,23R lactone > 25R-OHD3-26,23S lactone = in vivo 25-OHD3-26,23 lactone. When tested in a high-performance liquid chromatographic system which separates all four stereoisomers, the in vivo 25-OHD3-26,23 lactone comigrated with synthetic 25R-OHD3-26,23S lactone. These data firmly establish that the naturally-occurring 25-OHD3-26,23 lactone has the 25R, 23S stereochemistry. In addition, these data are the first to demonstrate that the four stereoisomers of 25-OHD3-26,23 lactone have different affinities for the plasma vitamin D binding protein and the 1,25-(OH)2D cytosol receptor.  相似文献   

3.
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   

4.
Kidney homogenates from chicks fed a vitamin D-deficient diet for 10 days and supplemented with 6.5 nmol of vitamin D3 48 hr prior to sacrifice metabolized invitro [3H]-25-hydroxyvitamin D3 (25-OH-D3) to 24,25-dihydroxyvitamin D3 [24,25-(OH)2-D3] and 3 other metabolites (peaks A, C and E). When the homogenates were incubated with purified [3H]-24,25-(OH)2-D3, 3 similar metabolites (peaks A′, C′ and E′) were produced. On high pressure liquid chromatography, peaks A, C and E migrated to exactly the same respective positions as peaks A′, C′ and E′. Kidney homogenates from D-deficient chicks failed to produce these metabolites from [3H]-25-OH-D3 or [3H]-24,25-(OH)2-D3. These results strongly suggest that the new metabolites reported here are synthesized via 24,25-(OH)2-D3 in the kidney of chicks supplemented with vitamin D3.  相似文献   

5.
The action of parathyroid extract (PTE) on the renal metabolism of 25-hydroxyvitamin D3 (25-OHD3) was evaluated in rat models for strontium rickets and hypoparathyroidism. PTE elevated the production of 1α,25-(OH)2D3 and suppressed the synthesis of 24,25-(OH)2D3 in both animal models. Part of strontium's action in suppressing 1α,25-(OH)2D3 and stimulating 24,25-(OH)2D3 synthesis in strontium rickets appears to involve a decrease in parathyroid hormone (PTH) secretion and/or action. Calcitonin (CT) was not implicated in the cation's action. Thyroparathyroidectomized rats showed a low level of 1α,25-(OH)2D3 production which increased four- to eightfold following chronic PTE treatment. PTH appears to be the major calcium regulatory hormone involved in modulation of renal 25-OHD3 metabolism.  相似文献   

6.
The T47D human breast cancer cell line contains a specific binding protein for 1.25-(OH)2D3, with 15000 sites per cell. The Kd (1.1 × 10?10 M) and sedimentation coefficient on sucrose gradients (3.7S) are the same as those reported for the 1,25-(OH)2D3 receptor in other tissues. Other vitamin D3 metabolites bound to the receptor with an order of affinities 1,25-(OH)2D3 > 1,24,25-(OH)3D3 > 25-OHD3 > 24,25-(OH)2D3 > D3. A new analogue 1β,25-(OH)2D3 was only as effective as 24,25-(OH)2D3 at displacing the hormone from the receptor. Cell growth was stimulated in a dose dependent manner by the addition of 1,25-(OH)2D3 (up to 0.8 nM) to the medium. A higher concentration of hormone was without effect.  相似文献   

7.
Plasma 1,25-dihydroxyvitamin D (1,25-(OH)2D) concentration was shown to decrease during bed rest in several studies when baseline plasma 25-hydroxyvitamin D (25-OHD) concentration was sub-optimal. Dahl salt-sensitive female (S) rats, but not Dahl salt-resistant female (R) rats, demonstrated a 50% decrease in plasma 1,25-dihydroxycholecalciferol (1,25-(OH)2D3) concentration after 28 days of hind limb unloading (HU, disuse model) during low salt intake (0.3%). We tested the vitamin D endocrine system response of female S rats to hind limb unloading during high salt intake (2%, twice that of standard rat chow to mimic salt intake in the USA). Hind limb unloading resulted in lower plasma 25-OHD3 concentrations in S-HU rats than in R-HU rats (P < 0.05) and greater urinary loss of 25-OHD3 by S-HU rats than by S rats (P < 0.05). Plasma 1,25-(OH)2D3 concentration of S-HU rats was half that of S rats, but was unchanged in R-HU rats. The association of low plasma 25-OHD concentration with decrease in plasma 1,25-(OH)2D concentration of hind limb unloaded rats and of bed rest participants (published studies) suggests that low vitamin D status might be a risk factor for decrease in plasma vitamin D hormone concentration during long-term immobilization or bed rest.  相似文献   

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.
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   

10.
Matrix metalloproteinases (MMPs) are a group of enzymes with the potential to degrade extracellular matrix proteins. One of the MMPs, stromelysin-1 (MMP-3) has been localized to extracellular matrix vesicles in growth plate chondrocyte cultures, suggesting involvement of this enzyme in remodeling of the extracellular matrix during endochondral development, a process which is regulated by the vitamin D metabolites, 1,25-(OH)2D3 and 24,25-(OH)2D3. To determine whether stromelysin-1 is regulated by vitamin D as well, confluent cultures of cells derived from growth zone (GC) and resting zone (RC) rat costochondral cartilage were treated with 1α,25-(OH)2D3 (1,25) and 24R,25-(OH)2D3 (24,25), respectively, and the effect on stromelysin-1 assessed by casein gel zymography and Western blots. Although stromelysin-1 activity was enriched in the matrix vesicle fraction, only the plasma membrane enzyme was affected by the treatment; 1,25 and 24,25 caused a marked decrease in plasma membrane stromelysin-1 activity in their target cells. Since plasma membrane protein kinase C (PKC) activity is stimulated by 1,25 and 24,25, we hypothesized that stromelysin-1 activity was regulated by the vitamin D metabolites via PKC-dependent phosphorylation. To test this, membrane fractions (containing endogenous PKCα and ζ as well as stromelysin-1) were incubated in the presence of purified rat brain PKC and/or recombinant human (rh) stromelysin-1 and [γ32P]-ATP and anti-stromelysin-1 immunoprecipitates were analyzed by autoradiography and Western blots. Immuno-phospho-stromelysin-1 was localized to a 52-kDa band in the plasma membrane fraction only; no phosphorylation was observed in the matrix vesicle fraction. Selective inhibitors of PKC activity demonstrated that phosphorylation was inhibited by H7 and low concentrations of H8, but not by HA1004, indicating that PKC, not PKA, was responsible. Protein phosphatase 2A, (PP2A), a serine/threonine-specific phosphatase, selectively removed the radiolabel in a time-dependent manner, providing further support for a PKC-dependent phosphorylation mechanism. Incubation of resting zone cell plasma membranes with 24,25, but not 1,25, resulted in phosphorylation of stromelysin-1, demonstrating that the nongenomic effect was metabolite-specific. This suggests that this may be one mechanism by which vitamin D metabolites regulate stromelysin-1 activity and that PKC-dependent phosphorylation inhibits the metalloproteinase. © 1996 Wiley-Liss, Inc.  相似文献   

11.
Vitamin D3 (cholecalciferol) is endogenously produced in the skin of primates when exposed to the appropriate wavelengths of ultraviolet light (UV-B). Common marmosets (Callithrix jacchus) maintained indoors require dietary provision of vitamin D3 due to lack of sunlight exposure. The minimum dietary vitamin D3 requirement and the maximum amount of vitamin D3 that can be metabolized by marmosets is unknown. Observations of metabolic bone disease and gastrointestinal malabsorption have led to wide variation in dietary vitamin D3 provision amongst research institutions, with resulting variation in circulating 25-hydroxyvitamin D3 (25(OH)D3), the accepted marker for vitamin D sufficiency/deficiency. Multiple studies have reported serum 25(OH)D3 in captive marmosets, but 25(OH)D3 is not the final product of vitamin D3 metabolism. In addition to serum 25(OH)D3, we measured the most physiologically active metabolite, 1,25-dihydroxyvitamin D3 (1,25(OH)2D3), and the less well understood metabolite, 24,25-dihydroxyvitamin D3 (24,25(OH)2D3) to characterize the marmoset's ability to metabolize dietary vitamin D3. We present vitamin D3 metabolite and related serum chemistry value colony reference ranges in marmosets provided diets with 26,367 (Colony A, N = 113) or 8,888 (Colony B, N = 52) international units (IU) of dietary vitamin D3 per kilogram of dry matter. Colony A marmosets had higher serum 25(OH)D3 (426 ng/ml [SD 200] vs. 215 ng/ml [SD 113]) and 24,25(OH)2D3 (53 ng/ml [SD 35] vs. 7 ng/ml [SD 5]). There was no difference in serum 1,25(OH)2D3 between the colonies. Serum 1,25(OH)2D3 increased and 25(OH)D3 decreased with age, but the effect was weak. Marmosets tightly regulate metabolism of dietary vitamin D3 into the active metabolite 1,25(OH)2D3; excess 25(OH)D3 is metabolized into 24,25(OH)2D3. This ability explains the tolerance of high levels of dietary vitamin D3 by marmosets, however, our data suggest that these high dietary levels are not required.  相似文献   

12.
Cultured osteogenic sarcoma (OS) cells have been used here to study the internalization of 1,25(OH)2D3 and other major metabolites of D3 by cells. Intact OS cells incubated for 1h at 37°C in medium containing [3H]1,25(OH)2D3 at low concentrations (0.16 to 1.6nM) take up and retain this hormone with high affinity (Kd=3.3×10?10M) similar to that found for the hormone-receptor interaction in cytosol preparations. Vitamin D3 and its major metabolites such as 25(OH)D3 or 24,25(OH)2D3, even at supraphysiological concentrations, are not internalized by the cells when small amounts of plasma D binding protein (DBP) or human alpha-globulin are added to the incubation medium. This phenomenon can be exploited to develop a non-chromatographic cytoreceptor assay for 1,25(OH)2D3.  相似文献   

13.
Serum and post-microsomal supernatants of human lymphocyte, erythrocyte, skeletal muscle and parathyroid adenoma homogenates were examined for specific binding of 25-hydroxycholecalciferol (25-OHD3) and 1,25-dihydroxycholecalciferol (1,25-(OH)2D3). Muscle, lymphocytes and parathyroid adenomata extracts contained a 6-S 25-OHD3-binding protein which was not found in erythrocyte extracts, and which was distinct from the smaller serum transport α-globulin. A cathodal, 1,25-(OH)2D3-binding protein, which sedimented at 3–4 S was also detected in parathyroid tissue. These observations suggest the possibility of direct physiologic interaction between vitamin D metabolites and nucleated human tissues other than intestine and bone.  相似文献   

14.
1,25-(OH)2D3 (1,25) and 24,25-(OH)2D3 (24,25) 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 analogues of 1,25 that have been modified on the A-ring (2a, 2b) and are only 0.1% as effective in binding to the VDR as 1,25, to examine the role of the VDR in the response of rat costochondral resting zone (RC) and growth zone (GC) chondrocytes to 1,25 and 24,25. Chondrocyte proliferation ([3H]-thymidine incorporation), proteoglycan production ([35S]-sulfate incorporation), and second messenger activation (activity of protein kinase C) were measured after treatment with 10-8 M 1,25, 10-7 M 24,25, or the analogues at 10-9–10-6 M. Both analogues inhibited proliferation of both cell types, as did 1,25 and 24,25. Neither 2a nor 2b had an effect on proteoglycan production by GCs or RCs. 2a caused a dose-dependent stimulation of protein kinase C (PKC) that was not inhibited by cycloheximide or actinomycin D in either GC or RC cells. 2b, on the other hand, had no effect on PKC activity in RCs and only a slight stimulatory effect in GCs. Both cells produce matrix vesicles, extracellular organelles associated with the initial stages of calcification, in culture that are regulated by vitamin D metabolites. Since these organelles contain no DNA or RNA, they provide an excellent model for studying the mechanisms used by vitamin D metabolites to mediate their nongenomic effects. When matrix vesicles were isolated from naive cultures of growth zone cells and treated with 2a, a dose-dependent inhibition of PKC activity was observed that was similar to that found with 1,25-(OH)2D3. Plasma membranes contained increased PKC activity after treatment with 2a, but the magnitude of the effect was less than that seen with 1,25-(OH)2D3. Analogue 2b had no affect on PKC activity in either membrane fraction. When matrix vesicles from resting zone chondrocyte cultures were treated with 24,25-(OH)2D3, a significant decrease in PKC activity was observed. No change in enzyme activity was found for either 1,25-(OH)2D3 or the analogues. PKC activity in the plasma membrane fraction, however, was increased by 24,25-(OH)2D3 as well as by analogue 2a. This study shows that these analogues, with little or no binding to the vitamin D receptor, can affect cell proliferation and PKC activity, but not proteoglycan production. The direct membrane effect is analogue specific and cell maturation dependent. Further, by eliminating the VDR-mediated component of the cellular response, we have provided further evidence for the existence of a membrane receptor(s) involved in mediating nongenomic effects of vitamin D metabolites. J. Cell. Physiol. 171:357–367, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

15.
We investigated the occurrence of rickets in adolescent tamarins (Saguinus imperator) residing at the Los Angeles Zoo. Compared to tamarins in the same colony without clinical evidence of bone disease (N = 6), rachitic platyrrhines (N = 3) had a decrease in their serum calcium concentration (P < .05). The affected tamarins also had lower serum 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3) levels than did nonaffected colony mates, but 2–10-fold higher concentrations than in Old World primates of a comparable developmental stage. New World primates in many different genera are known to exhibit target organ resistance to the active vitamin D3 metabolite, 1,25-(OH)2D3, compensated by maintenance of high circulating concentrations of 1,25-(OH)2D3. The relatively low serum 1,25-(OH)2D3 concentration in rachitic tamarins and ultraviolet B radiation deficient environment of these primates suggested that bone disease may be linked to a deficiency in substrate for 1,25-(OH)2D3, 25 hydroxyvtamin D3 (25-OHD3). Chronic exposure of platyrrhines in three different vitamin D resistant genera to an artificial UVB source resulted in 1) a significant increase in the mean serum 25-OHD3 (P < .001) and 1,25-(OH)2D3 (P < .02) level over that encountered in platyrrhines not exposed to UVB; and 2) prevention of rachitic bone disease in irradiated individuals. These data further show that the serum 25-OHD3 and 1,25-OH2D3 levels are positively correlated in vitamin D-resistant platyrrhines (r = 0.64; P= .0014) and suggest that a compromise in cutaneous vitamin D3 production by means of UVB deprivation may limit necessary 1,25-(OH)2D3 production. © 1992 Wiley-Liss, Inc.  相似文献   

16.
The yolk sac of the pregnant rat which functions as a true placenta is a target organ for vitamin D. This tissue can hydroxylate in position 24 both 25-hydroxy- and 1,25-dihydroxyvitamin D3 (25-OHD3 and 1,25-(OH)2D3). The present report describes an in vitro model for the study of 1,25-(OH)2D3 action on the further metabolism of 25-OH[3H]D3 and 1,25-(OH)2[3H]D3 by yolk sac. The tissue explants were preincubated with 1,25-(OH)2D3 for 18 h in a serum-free culture medium. Physiological concentrations of 1,25-(OH)2D3 were the most effective in stimulating (7.5-fold) the 1,25-(OH)2D3 24-hydroxylase, while the 25-OHD3 24-hydroxylase stimulation (4-fold) required a 1,25-(OH)2D3 concentration of 10(-7) M. The stimulating effect of 1,25-(OH)2D3 on the 1,25-(OH)2D3 24-hydroxylase was temperature-dependent, and, since its was inhibited by actinomycin D and cycloheximide, required de novo protein synthesis. 1,24,25-(OH)3D3, 25-OHD3, and 24,25-(OH)2D3 were 10- to 1000-fold less potent than 1,25-(OH)2D3 in inducing the 1,25-(OH)2D3 hydroxylase. Our results strongly suggest that 1,25-(OH)2D3 regulated the 1,25-(OH)2D3 24-hydroxylase by a receptor-mediated process. Furthermore, 1,25-(OH)2D3 at 10(-9) M induced within 4 h an increase of its own degradation and the formation of an as yet unidentified major 1,25-(OH)2[3H]D3 metabolite. We conclude that the yolk sac can participate in the regulation of 1,25-(OH)2D3 concentration in the fetoplacental unit.  相似文献   

17.
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   

18.
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   

19.
AimsLigands for the vitamin D receptor (VDR) regulate apolipoprotein A-I (apo A-I) gene expression in a tissue-specific manner. The vitamin D metabolite 24, 25-dihydroxycholecalciferol (24, 25-(OH)2D3) has been shown to possess unique biological effects. To determine if 24, 25-(OH)2D3 modulates apo A-I gene expression, HepG2 hepatocytes and Caco-2 intestinal cells were treated with 24, 25-(OH)2D3 or its precursor 25-OHD3.Main methodsApo A-I protein levels and mRNA levels were measured by Western and Northern blotting, respectively. Changes in apo A-I promoter activity were measured using the chlorampenicol acetytransferase assay.Key findingsTreatment with 24, 25-(OH)2D3, but not 25-OHD3, inhibited apo A-I secretion in HepG2 and Caco-2 cells and apo A-I mRNA levels and apo A-I promoter activity in HepG2 cells. To determine if 24, 25-(OH)2D3 represses apo A-I gene expression through site A, the nuclear receptor binding element that is essential for VDRs effects on apo A-I gene expression, HepG2 cells were transfected with plasmids containing or lacking site A. While the site A-containing plasmid was suppressed by 24, 25-(OH)2D3, the plasmid lacking site A was not. Likewise, treatment with 24, 25-(OH)2D3 suppressed reporter gene expression in cells transfected with a plasmid containing site A in front of a heterologous promoter. Finally, antisense-mediated VDR depletion failed to reverse the silencing effects of 24, 25-(OH)2D3 on apo A-I expression.SignificanceThese results suggest that the vitamin D metabolite 24, 25-(OH)2D3 is an endogenous regulator of apo A-I synthesis through a VDR-independent signaling mechanism.  相似文献   

20.
Serum concentrations of 25-hydroxycholecalciferol (25-OHD), 24,25-dihydroxycholecalciferol (24,25-(OH)2D), and 1,25-dihydroxycholecalciferol (1,25-(OH)2D) were measured at monthly intervals throughout the year in eight normal subjects. 25-OHD was measured by competitive protein-binding assay after Sephadex LH 20 chromatography, 24,25-(OH)2D by competitive protein-binding assay after Sephadex LH 20 and high-pressure chromatography, and 1,25-(OH)2D by radioimmunoassay after the same separation procedure as for 24,25-(OH)2D. A seasonal variation, apparently dependent on exposure to ultraviolet light, was found for all three metabolites. A study in six other normal subjects showed that there was no diurnal rhythm in any of the metabolites. Oral administration of 2 microgram 1,25-(OH)2D caused a sharp rise in serum concentrations of 1,25-(OH)2D and no change in the concentrations of the two other metabolites, but by 12 hours the 1,25-(OH)2D concentration had returned to the basal value. The concentrations of all three metabolites studied vary according to the season. Thus to interpret these concentrations in any subject the normal range for the particular season must be referred to.  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