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.
Monocytic differentiation-inducing activity of 26,26,26,27,27,27-hexafluoro-1 alpha,25-dihydroxyvitamin D3 [26,27-F6-1 alpha,25-(OH)2D3] was re-evaluated in human promyelocytic leukemia (HL-60) cells in serum-supplemented or serum-free culture. The order of in vitro potency for reducing nitroblue tetrazolium (NBT) was 26,27-F6-1 alpha,25-(OH)2D3 greater than 1 alpha, 25-dihydroxyvitamin D3 [1 alpha,25-(OH)2D3] = 26,26,26,27,27,27-F6-1 alpha,23(S), 25-trihydroxyvitamin D3 [26,27-F6-1 alpha,23(S), 25-(OH)3D3] under serum-supplemented culture conditions, whereas the order was 1 alpha, 25-(OH)2D3 = 26,27-F6-1 alpha,25-(OH)2D3 greater than 26,27-F6-1 alpha,23(S), 25-(OH)3D3 under serum-free culture conditions. This rank order for differentiation-inducing activity under serum-free culture conditions correlated well with the binding affinity of these analogs for vitamin D3 receptor of HL-60 cells. The order of relative % binding affinity for the vitamin D-binding protein in fetal calf serum was 1 alpha,25-(OH)2D3 (100%) much greater than 26,27-F6-1 alpha,25-(OH)2D3 (5.1%) greater than 26,27-F6-1 alpha,23(S), 25-(OH)3D3 (less than 1%). These results suggest that serum vitamin D-binding proteins apparently modulate monocytic differentiation of HL-60 cells by 26,27-F6-1 alpha,25-(OH)2D3 under serum-supplemented culture conditions.  相似文献   

2.
A Honda  Y Mori  S Otomo  S Ishizuka  N Ikekawa 《Steroids》1991,56(3):142-147
Monocytic differentiation-inducing activity of steroidal side chain-lengthened 26,27-dialkyl analogs of 1 alpha,25-dihydroxyvitamin D3 was examined in human promyelocytic leukemia (HL-60) cells in serum-supplemented or serum-free culture. The order of in vitro potency for reducing nitroblue tetrazolium was 1 alpha,25-dihydroxy-26,27-dimethylvitamin D3 greater than or equal to 1 alpha,25-dihydroxy-26,27-diethylvitamin D3 much greater than 1 alpha,25-dihydroxy-26,27-dipropylvitamin D3 under serum-free culture conditions. Analysis by sucrose density-gradient centrifugation or polyethylene glycol precipitation technique showed that the potency order for differentiation-inducing activity correlated well with binding affinity of these analogs for vitamin D3 receptor of HL-60 cells. Under serum-supplemented culture conditions, the lack of correlation between biologic activity and analog-binding affinity for receptor was caused by differences in binding affinity of these analogs for serum vitamin D-binding proteins. These results suggest that serum vitamin D-binding proteins apparently modulate monocytic differentiation of HL-60 cells by these analogs under serum-supplemented culture conditions.  相似文献   

3.
J K Addo  N Swamy  R Ray 《Steroids》1999,64(4):273-282
In this article, we describe the development of a general synthetic strategy to functionalize the C-6 position of vitamin D3 and its biologically important metabolites, i.e. 25-hydroxyvitamin D3 (25-OH-D3) and 1alpha,25-dihydroxyvitamin D3 [1,25(OH)2D3]. We employed Mazur's cyclovitamin D method to synthesize vitamin D3 analogs with several functionalities at the C-6 position. In addition, we synthesized 6-(3-hydroxypropyl) and 6-[(2-bromoacetoxy)propyl] derivatives of 25-OH-D3 15 and 16, respectively, and 6-(3-hydroxypropyl) derivative of 1,25(OH)2D3 17. Competitive binding assays of 15-17 with human serum vitamin D-binding protein showed that all these analogs specifically bound to this protein, although with significantly lower affinity than the 25-OH-D3, the strongest natural binder, but with comparable affinity with 1,25(OH)2D3, the hormone. On the other hand, 6-[3-hydroxypropyl], 1alpha,25-dihydroxyvitamin D3 17 did not show any specific binding for recombinant nuclear vitamin D receptor. These results indicated that the region containing the C-6 position of the parent seco-steroid [1,25(OH)2D3] may be an important recognition marker towards vitamin D receptor binding. Information, delineated in this article, will be important for evaluating structure-activity relationship in synthetic analogs of vitamin D and its metabolites.  相似文献   

4.
5.
Our recent studies with 2-(3'-hydroxypropylidene) analogs of 1alpha,25-dihydroxy-19-norvitamin D(3) showed that this 2-substituent creates compounds with very potent biological activity. In the continuing search for vitamin D compounds with selective activity profiles, we prepared a series of 1alpha-hydroxy-19-norvitamin D analogs characterized by the presence of a 3'-hydroxypropylidene substituent at C-2 and a truncated side chain. These vitamin D compounds were efficiently prepared using convergent syntheses. The C,D-fragments, namely the Grundmann ketones 19, 20, 27, 36 and 37 were synthesized from the known 8beta-benzoyloxy-22-aldehydes 12 and 29. These hydrindanones were subjected to Lythgoe type Wittig-Horner coupling with phosphine oxide 21, prepared by us previously, and after hydroxyl deprotection the set of 19-norvitamins 7-11 was successfully obtained. According to our expectations, all analogs (with an exception of the 20R-compound 7) have pronounced in vitro activity. When compared to the natural hormone 1alpha,25-(OH)(2)D(3) (1), they show the same or only slightly reduced affinity for the vitamin D receptor while being similarly effective as 1 in differentiation of HL-60 cells into monocytes.  相似文献   

6.
The multiple physiological properties of vitamin D-binding protein (DBP) include organ-specific transportation of vitamin D(3) and its metabolites, manifested by its ability to bind vitamin D sterols with high affinity. In the present investigation we probed the vitamin D sterol-binding pocket of human DBP with affinity labeling analogs of 25-hydroxyvitamin D(3) ?25-OH-D(3) and 1, 25-dihydroxyvitamin D(3) ?1,25(OH)(2)D(3) containing bromoacetate alkylating probe at C-3 (A-ring), C-6 (triene), C-11 (C-ring), and C-19 (exocyclic methylene) of the parent sterol. Competitive binding assays with DBP showed approximately 22-, 68-, and 2000-fold decrease in the binding of 1,25(OH)(2)-D(3)-11-BE, 25-OH-D(3)-3-BE, and 25-OH-D(3)-6-BE, respectively, compared to that seen with 25-OH-D(3), while there was no significant difference in the DBP-binding affinity of 25-OH-D(3)-19-BE and 25-OH-D(3). Surprisingly, ?(14)C25-OH-D(3)-11-BE and ?(14)C1, 25(OH)(2)-D(3)-19-BE failed to label DBP despite high-affinity DBP-binding, indicating the absence of any nucleophilic amino acid in the vicinity of their bromoacetate moiety to form a covalent bond, while these analogs are inside the binding pocket. In contrast, ?(14)C25-OH-D(3)-6-BE and ?(14)C25-OH-D(3)-3-BE specifically labeled DBP. BNPS-skatole digestion of DBP labeled with ?(14)C25-OH-D(3)-3-BE or ?(14)C25-OH-D(3)-6-BE produced two peptides (M(r) 17,400 and 33,840), with radioactivity associated with the N- and C-terminal peptides, respectively, raising the possibility that either different areas of the same vitamin D sterol-binding pocket, or different domains of DBP might be labeled by these analogs. Successful affinity labeling of recombinant domain I (1-203) of DBP with both reagents indicated that different areas of the same vitamin D-binding pocket (domain I) were labeled. These affinity analogs are potentially useful for "mapping" the vitamin D sterol-binding pocket and developing a functional model.  相似文献   

7.
Vitamin D compounds added to the culture medium induce differentiation of human myeloid leukemia cells (HL-60 cells) by binding to a specific cytosol receptor protein. This system provides a biologically relevant and technically simple assay to examine the relationship between molecular structure and biological activity of vitamin D compounds. Using this culture system, the biological activity of 24,24-F2-1 alpha,25(OH)2D3 and 1 alpha,25(OH)2D3-26,23-lactone was assayed. 24,24-F2-1 alpha,25(OH)2D3 was four to seven times more potent than 1 alpha,25(OH)2D3 in inducing phagocytosis and C3 rosette formation of HL-60 cells, though both compounds bound equally well to the cytosol receptor, suggesting that the defuorination at the 24-carbon position may stimulate membrane permeability of the compound. 1 alpha,25(OH)2D3-26,23-lactone, on the other hand, was only 1/200th as active as 1 alpha,25(OH)2D3. The binding affinity of the lactone for the cytosol receptor was identical with that of 1 alpha (OH)D3, suggesting that the lactone formation between the 26 and 23 positions masks the function of the 25-hydroxyl group. The binding affinity of vitamin D3 derivatives to the specific cytosol receptor of HL-60 cells was well correlated with that of intestinal cytosol protein specifically bound to 1 alpha,25(OH)2D3.  相似文献   

8.
Analogs of 1alpha,25-dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)] activate both genomic mechanisms via the nuclear vitamin D(3) receptor (nVDR) and nongenomic pathways via the plasma membrane vitamin D(3) receptor (pmVDR). Both of these pathways are normally activated by 1alpha,25(OH)(2)D(3), but as a result of synthesis of numerous analogs of 1alpha,25(OH)(2)D(3) these pathways can be distinguished. We used increasing doses of vitamin D(3) analogs to determine their potencies of action on these two distinct pathways, measuring calcium channel potentiation as an indicator of the nongenomic action and measuring increases in osteocalcin mRNA and protein release and bone resorption as indicators of genomic action. We found that both 25(OH)-16,23E-diene-D(3) (R) and 1alpha,25(OH)(2)-16,23E-diene-D(3) (A) are 10-fold more potent than 1alpha,25(OH)(2)D(3) for activation of the nongenomic pathway because double bonds in the side chain and the D ring increase the affinity for calcium channel potentiation. While the C-1alpha-hydroxyl group is not necessary for potentiation of calcium channels, methyl groups at this position can alter the affinity for calcium channel potentiation. On the other hand, 1000 fold higher concentrations of nongenomic analogs were needed compared to 1alpha,25(OH)(2)D(3) to increase osteocalcin mRNA or protein release. 1alpha,25-Dihydroxy-16-ene-23-yne-26,27-hexafluorovitamin D(3), (E) is an agent that is 10 fold more potent than 1alpha,25(OH)(2)D(3) at increasing osteocalcin mRNA and protein release, whereas 1alpha,25(OH)(2)-3-epi-D(3) increases osteocalcin mRNA and protein with a potency over 10 fold lower than 1alpha,25(OH)(2)D(3). These results suggest that double bonds in the side chain and the D ring stabilize action on the nongenomic pathway whereas F(6) on the terminal portion of the side chain increases potency for nVDR. On the other hand, while the C-1alpha-hydroxyl group is necessary for activation of genomic events via nVDR, the activation of nongenomic events occurs in the absence of this group.  相似文献   

9.
10.
We examined the effects of two novel 1alpha,25-dihydroxyvitamin D(3)-26,23-lactone (1alpha,25-(OH)(2)D(3)-26,23-lactone) analogs on 1alpha,25(OH)(2)D(3)-induced differentiation of human leukemia HL-60 cells thought to be mediated by the genomic action of 1alpha, 25-dihydroxyvitamin D(3) (1alpha,25-(OH)(2)D(3)) and of acute promyelocytic leukemia NB4 cells thought to be mediated by non-genomic actions of 1alpha,25-(OH)(2)D(3). We found that the 1alpha,25-(OH)(2)D(3)-26,23-lactone analogs, (23S)-25-dehydro-1alpha-hydroxyvitamin D(3)-26,23-lactone (TEI-9647) and (23R)-25-dehydro-1alpha-hydroxyvitamin D(3)-26,23-lactone (TEI-9648), inhibited differentiation of HL-60 cells induced by 1alpha,25-(OH)(2)D(3). However, 1beta-hydroxyl diastereomers of these analogs, i.e. (23S)-25-dehydro-1beta-hydroxyvitamin D(3)-26, 23-lactone (1beta-TEI-9647) and (23R)-25-dehydro-1beta-hydroxyvitamin D(3)-26,23-lactone (1beta-TEI-9648), did not inhibit differentiation of HL-60 cells caused by 1alpha,25-(OH)(2)D(3). A separate study showed that the nuclear vitamin D receptor (VDR) binding affinities of the 1-hydroxyl diastereomers were about 200 and 90 times weaker than that of 1alpha-hydroxyl diastereomers, respectively. Moreover, none of these lactone analogs inhibited NB4 cell differentiation induced by 1alpha,25-(OH)(2)D(3). In contrast, 1beta,25-dihydroxyvitamin D(3) (1beta,25-(OH)(2)D(3)) and 1beta,24R-dihydroxyvitamin D(3) (1beta,24R-(OH)(2)D(3)) inhibited NB4 cell differentiation but not HL-60 cell differentiation. Collectively, the results suggested that 1-hydroxyl lactone analogs, i.e. TEI-9647 and TEI-9648, are antagonists of 1alpha,25-(OH)(2)D(3), specifically for the nuclear VDR-mediated genomic actions, but not for non-genomic actions.  相似文献   

11.
The secosteroid hormone, 1alpha,25-dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)], induces differentiation of the human promyelocytic leukemia (HL-60) cells into monocytes/macrophages. At present, the metabolic pathways of 1alpha,25(OH)(2)D(3) and the biologic activity of its various natural intermediary metabolites in HL-60 cells are not fully understood. 1alpha,25(OH)(2)D(3) is metabolized in its target tissues via modifications of both the side chain and the A-ring. The C-24 oxidation pathway, the main side chain modification pathway initiated by hydroxylation at C-24 leads to the formation of the end product, calcitroic acid. The C-23 and C-26 oxidation pathways, the minor side chain modification pathways initiated by hydroxylations at C-23 and C-26 respectively together lead to the formation of the end product, 1alpha,25(OH)(2)D(3)-lactone. The C-3 epimerization pathway, the newly discovered A-ring modification pathway is initiated by epimerization of the hydroxyl group at C-3 to form 1alpha,25-dihydroxy-3-epi-vitamin-D(3). We performed the present study first to examine in detail the metabolism of 1alpha,25(OH)(2)D(3) in HL-60 cells and then to assess the ability of the various natural intermediary metabolites of 1alpha,25(OH)(2)D(3) in inducing differentiation and in inhibiting clonal growth of HL-60 cells. We incubated HL-60 cells with [1beta-(3)H] 1alpha,25(OH)(2)D(3) and demonstrated that these cells metabolize 1alpha,25(OH)(2)D(3) mainly via the C-24 oxidation pathway and to a lesser extent via the C-23 oxidation pathway, but not via the C-3-epimerization pathway. Three of the natural intermediary metabolites of 1alpha,25(OH)(2)D(3) derived via the C-24 oxidation pathway namely, 1alpha,24(R),25-trihydroxyvitamin D(3), 1alpha,25-dihydroxy-24-oxovitamin D(3) and 1alpha,23(S),25-trihydroxy-24-oxovitamin D(3) [1alpha,23(S),25(OH)(3)-24-oxo-D(3)] were almost as potent as 1alpha,25(OH)(2)D(3) in terms of their ability to differentiate HL-60 cells into monocytes/macrophages. We then selected 1alpha,23(S),25(OH)(3)-24-oxo-D(3) which has the least calcemic activity among all the three aforementioned natural intermediary metabolites of 1alpha,25(OH)(2)D(3) to examine further its effects on these cells. Our results indicated that 1alpha,23(S),25(OH)(3)-24-oxo-D(3) was also equipotent to its parent in inhibiting clonal growth of HL-60 cells and in inducing expression of CD11b protein. In summary, we report that 1alpha,25(OH)(2)D(3) is metabolized in HL-60 cells into several intermediary metabolites derived via both the C-24 and C-23 oxidation pathways but not via the C-3 epimerization pathway. Some of the intermediary metabolites derived via the C-24 oxidation pathway are found to be almost equipotent to 1alpha,25(OH)(2)D(3) in modulating growth and differentiation of HL-60 cells. In a previous study, the same metabolites when compared to 1alpha,25(OH)(2)D(3) were found to be less calcemic. Thus, the findings of our study suggest that some of the natural metabolites of 1alpha,25(OH)(2)D(3) may be responsible for the final expression of the noncalcemic actions that are presently being attributed to their parent, 1alpha,25(OH)(2)D(3).  相似文献   

12.
The secosteroid hormone 1alpha,25-dihydroxyvitamin D3 [1alpha,25(OH)2D3] is metabolized in its target tissues through modifications of both the side chain and the A-ring. The C-24 oxidation pathway, the previously well established main side chain modification pathway, is initiated by hydroxylation at C-24 of the side chain. The C-3 epimerization pathway, the newly discovered A-ring modification pathway, is initiated by epimerization of the hydroxyl group at C-3 of the A-ring. The end products of the metabolism of 1alpha,25(OH)2D3 through the C-24 oxidation and the C-3 epimerization pathways are calcitroic acid and 1alpha,25-dihydroxy-3-epi-vitamin-D3 respectively. During the past two decades, numerous noncalcemic analogs of 1alpha,25(OH)2D3 were synthesized. Several of the analogs have altered side chain structures and as a result some of these analogs have been shown to resist their metabolism through side chain modifications. For example, two of the analogs, namely, 1alpha,25-dihydroxy-16-ene-23-yne-vitamin D3 [1alpha,25(OH)2-16-ene-23-yne-D3] and 1alpha,25-dihydroxy-16-ene-23-yne-20-epi-vitamin D3 [1alpha,25(OH)2-16-ene-23-yne-20-epi-D3], have been shown to resist their metabolism through the C-24 oxidation pathway. However, the possibility of the metabolism of these two analogs through the C-3 epimerization pathway has not been studied. Therefore, in our present study, we investigated the metabolism of these two analogs in rat osteosarcoma cells (UMR 106) which are known to express the C-3 epimerization pathway. The results of our study indicate that both analogs [1alpha,25(OH)2-16-ene-23-yne-D3 and 1alpha,25(OH)2-16-ene-23-yne-20-epi-D3] are metabolized through the C-3 epimerization pathway in UMR 106 cells. The identity of the C-3 epimer of 1alpha,25(OH)2-16-ene-23-yne-D3 [1alpha,25(OH)2-16-ene-23-yne-3-epi-D3] was confirmed by GC/MS analysis and its comigration with synthetic 1alpha,25(OH)2-16-ene-23-yne-3-epi-D3 on both straight and reverse-phase HPLC systems. The identity of the C-3 epimer of 1alpha,25(OH)2-16-ene-23-yne-20-epi-D3 [1alpha,25(OH)2-16-ene-23-yne-20-epi-3-epi-D3] was confirmed by GC/MS and 1H NMR analysis. Thus, we indicate that vitamin D analogs which resist their metabolism through the C-24 oxidation pathway, have the potential to be metabolized through the C-3 epimerization pathway. In our present study, we also noted that the rate of C-3 epimerization of 1alpha,25(OH)2-16-ene-23-yne-20-epi-D3 is about 10 times greater than the rate of C-3 epimerization of 1alpha,25(OH)2-16-ene-23-yne-D3. Thus, we indicate for the first time that certain structural modifications of the side chain such as 20-epi modification can alter significantly the rate of C-3 epimerization of vitamin D compounds.  相似文献   

13.
14.
We have demonstrated that 1alpha,25-dihydroxyvitamin D(3)-26, 23-lactone analogs, (23S)- and (23R)-25-dehydro-1alpha-hydroxyvitamin D(3)-26,23-lactone (TEI-9647, TEI-9648, respectively), inhibit HL-60 cell differentiation induced by 1alpha,25-dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)], but not differentiation caused by all-trans retinoic acid (D. Miura et al., 1999, J. Biol. Chem. 274, 16392). To assess whether the antagonistic actions of TEI-9647 and TEI-9648 in HL-60 cells are related to 1alpha,25(OH)(2)D(3) breakdown, we investigated their effects on catabolism of 1alpha,25(OH)(2)D(3). In HL-60 cells, the C-24 but not the C-23 side-chain oxidation pathway of 1alpha,25(OH)(2)D(3) has been reported. Here we demonstrate that 1alpha,25(OH)(2)D(3) was metabolized both to 24,25,26,27-tetranor-1alpha,23-(OH)(2)D(3) and 1alpha,25(OH)(2)D(3)-26,23-lactone; thus HL-60 cells constitutively possess both the 24- and the 23-hydroxylases. Metabolism of 1alpha, 25(OH)(2)D(3) was strongly suppressed by 10(-7) M TEI-9647 or 10(-6) M TEI-9648. 1alpha,25(OH)(2)D(3) alone slightly induced 24-hydroxylase gene expression by 8 h with full enhancement by 24-48 h; this induction was inhibited by 10(-6) M TEI-9647 and 10(-6) M TEI-9648 (86.2 and 31.9%, respectively) 24 h after treatment. However, analogs of TEI-9647 and TEI-9648 without the 25-dehydro functionality induced 24-hydroxylase gene expression. These results indicate that TEI-9647 and TEI-9648 clearly mediate their stereoselective antagonistic actions independent of their actions to block the catabolism of 1alpha,25(OH)(2)D(3). Therefore, TEI-9647 and TEI-9648 appear to be the first antagonists specific for the nuclear 1alpha,25(OH)(2)D(3) receptor-mediated genomic actions of 1alpha,25(OH)(2)D(3) in HL-60 cells.  相似文献   

15.
1,25-Dihydroxyvitamin D3, the hormonal form of vitamin D, induces differentiation of HL-60 human promyelocytes into monocyte-like cells in vitro. We assessed the relative activity of 30 analogs of 1,25-dihydroxyvitamin D3 in inducing development of monocytic markers in HL-60 cells. The three differentiation markers assayed were nonspecific acid esterase activity, nitro blue tetrazolium reducing activity, and phagocytic capacity. Of the known metabolites of vitamin D, 1,25-dihydroxyvitamin D3 is the most active; 50% of the cells exhibit the mature phenotype following a 4-day treatment with 10(-8) M 1,25-dihydroxyvitamin D3. Removal of either the C-1 or C-25-hydroxyl group reduces activity by 2 orders of magnitude, while epimerization of the 1 alpha- to 1 beta-hydroxyl group virtually abolishes activity. Elongation of the steroidal side chain of 1,25-dihydroxyvitamin D3 by addition of one carbon at C-24 or C-26 improves the potency by an order of magnitude. Truncation of the steroidal side chain leads to a 10-fold reduction in activity for each carbon removed. Elimination of the C-26 and C-27 methyl groups reduces activity 100-fold. Analogs with short aliphatic side chains as 1 alpha-hydroxyhomo- and bishomopregnacholecalciferol have surprisingly high activity, being only 20-fold less potent than the natural hormone. The activity of most analogs in the HL-60 system parallels their known relative affinities for the well characterized 1,25-dihydroxyvitamin D3 receptor in chick intestine, providing further evidence that this function of 1,25-dihydroxyvitamin D3 is receptor mediated.  相似文献   

16.
1alpha,25-Dihydroxyvitamin D(3) (1alpha,25(OH)(2)D(3)) is known to inhibit the proliferation and invasiveness of prostate cancer cells. However, 1alpha,25(OH)(2)D(3) can cause hypercalcemia and is not suitable as a therapeutic agent. 19-Nor-vitamin D derivatives are known to be less calcemic when administered systemically. In order to develop more potent anti-cancer agents with less calcemic side effect, we therefore utilized (3)H-thymidine incorporation as an index for cell proliferation and examined the antiproliferative activities of nine C-2-substituted 19-nor-1alpha,25(OH)(2)D(3) analogs in the immortalized PZ-HPV-7 normal prostate cell line. Among the nine analogs we observed that the substitution with 2alpha- or 2beta-hydroxypropyl group produced two analogs having antiproliferative potency that is approximately 500- to 1000-fold higher than 1alpha,25(OH)(2)D(3). The (3)H-thymidine incorporation data were supported by the cell counting data after cells were treated with 1alpha,25(OH)(2)D(3), 19-nor-2alpha-(3-hydroxypropyl)-1alpha,25(OH)(2)D(3) or 19-nor-2beta-(3-hydroxypropyl)-1alpha,25(OH)(2)D(3) for 7 days. 19-Nor-2alpha-(3-hydroxypropyl)-1alpha,25(OH)(2)D(3) and 19-nor-2beta-(3-hydroxypropyl)-1alpha,25(OH)(2)D(3) were also shown to be about 10-fold more active than 1alpha,25(OH)(2)D(3) in cell invasion studies using prostate cancer cells. In conclusion, a substitution at the C-2 position of 19-nor-1alpha,25(OH)(2)D(3) molecule with a hydroxypropyl group greatly increased the antiproliferative and anti-invasion potencies. Thus, these two analogs could be developed to be effective therapeutic agents for treating early and late stages of prostate cancer.  相似文献   

17.
Vitamin D compounds added to the culture medium induce HL-60 cells to differentiate into macrophage/monocytes via a receptor mechanism. This system provides a biologically relevant assay for the study of biopotency of vitamin D analogs. Using this system, the biological activity of various fluorinated derivatives of vitamin D3 was compared with that of 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3). As assessed by cell morphology, nitroblue tetrazolium reduction and nonspecific esterase activity, 26,26,26,27,27,27-hexafluoro-1,25-dihydroxyvitamin D3 (26,27-F6-1,25-(OH)2D3) and 26,26,26,27,27,27-hexafluoro-1,24-dihydroxyvitamin D3 (26,27-F6-1,24-(OH)2D3) were about 10 times as potent as 1,25-(OH)2D3 in suppressing HL-60 cell proliferation and inducing cell differentiation. The biological activity of 26,26,26,27,27,27-hexafluoro-1-hydroxyvitamin D3 (26,27-F6-1-OH-D3) was equal to that of 1,25-(OH)2D3 in this system. 1,25-(OH)2D3 and its fluorinated analogs exerted their effects on HL-60 cells in a dose-dependent manner. HL-60 cells have a specific receptor for 1,25-(OH)2D3 with an apparent Kd of 0.25 nM, identical with that of chick intestinal receptor. While the binding affinities of 26,27-F6-1,25-(OH)2D3 and 26,27-F6-1,24-(OH)2D3 for chick intestinal receptor were lower than that of 1,25-(OH)2D3 by factors of 3 and 1.5, respectively, they were as competent as 1,25-(OH)2D3 in binding to HL-60 cell receptor. The ability of 26,27-F6-1-OH-D3 to compete for receptor protein from HL-60 cells and chick intestine was about 1/70 that of 1,25-(OH)2D3. These results indicate that trifluorination of carbons 26 and 27 of vitamin D3 can markedly enhance the effect on HL-60 cells.  相似文献   

18.
It is now well established that 1alpha,25(OH)2D3 is metabolized in its target tissues through the modifications of both side chain and A-ring. The C-24 oxidation pathway is the side chain modification pathway through which 1alpha,25(OH)2D3 is metabolized into calcitroic acid. The C-3 epimerization pathway is the A-ring modification pathway through which 1alpha,25(OH)2D3 is metabolized into 1alpha,25(OH)2-3-epi-D3. During the past two decades, a great number of vitamin D analogs were synthesized by altering the structure of both side chain and A-ring of 1alpha,25(OH)2D3 with the aim to generate novel vitamin D compounds that inhibit proliferation and induce differentiation of various types of normal and cancer cells without causing significant hypercalcemia. Previously, we used some of these analogs as molecular probes to examine how changes in 1alpha,25(OH)2D3 structure would affect its target tissue metabolism. Recently, several nonsteroidal analogs of 1alpha,25(OH)2D3 with unique biological activity profiles were synthesized. Two of the analogs, SL 117 and WU 515 lack the C-ring of the CD-ring skeleton of 1alpha,25(OH)2D3. SL 117 contains the same side chain as that of 1alpha,25(OH)2D3, while WU 515 contains an altered side chain with a 23-yne modification combined with hexafluorination at C-26 and C-27. Presently, it is unknown how the removal of C-ring from the CD-ring skeleton of 1alpha,25(OH)2D3 would affect its target tissue metabolism. In the present study, we compared the metabolic fate of SL 117 and WU 515 with that of 1alpha,25(OH)2D3 in both the isolated perfused rat kidney, which expresses only the C-24 oxidation pathway and rat osteosarcoma cells (UMR 106), which express both the C-24 oxidation and C-3 epimerization pathways. The results of our present study indicate that SL 117 is metabolized like 1alpha,25(OH)2D3, into polar metabolites via the C-24 oxidation pathway in both rat kidney and UMR 106 cells. As expected, WU 515 with altered side chain structure is not metabolized via the C-24 oxidation pathway. Unlike in rat kidney, both SL 117 and WU 515 are also metabolized into less polar metabolites in UMR 106 cells. These metabolites displayed GC and MS characteristics consistent with A-ring epimerization and were putatively assigned as C-3 epimers of SL 117 and WU 515. In summary, we report that removal of the C-ring from the CD-ring skeleton of 1alpha,25(OH)2D3 does not alter its target tissue metabolism significantly.  相似文献   

19.
20.
Construction of 25-hydroxy-steroidal side chain substituted with iodine at C-22 was elaborated on a model PTAD-protected steroidal 5,7-diene and applied to a synthesis of (22R)- and (22S)-22-iodo-1alpha,25-dihydroxyvitamin D3. Configuration at C-22 in the iodinated vitamins, obtained by nucleophilic substitution of the corresponding 22S-tosylates with sodium iodide, was determined by comparison of their iodine-displacement processes and cyclizations leading to isomeric five-membered (22,25)-epoxy-1alpha-hydroxyvitamin D3 compounds. Also, 20(22)-dehydrosteroids have been obtained and their structures established by 1H NMR spectroscopy. When compared to the natural hormone, (E)-20(22)-dehydro-1alpha,25-dihydroxyvitamin D3 was found 4 times less potent in binding to the porcine intestinal vitamin D receptor (VDR) and 2 times less effective in differentiation of HL-60 cells. 22-Iodinated vitamin D analogues showed somewhat lower in vitro activity, whereas (22,25)-epoxy analogues were inactive. Interestingly, it was established that (22S)-22-iodo-1alpha,25-dihydroxyvitamin D3 was 3 times more potent than its (22R)-isomer in binding to VDR and four times more effective in HL-60 cell differentiation assay. The restricted mobility of the side chain of both 22-iodinated vitamin D compounds was analyzed by a systematic conformational search indicating different spatial regions occupied by their 25-oxygen atoms. Preliminary data on the in vivo calcemic activity of the synthesized vitamin D analogues indicate that (E)-20(22)-dehydro-1alpha,25-dihydroxyvitamin D3 and 22-iodo-1alpha,25-dihydroxyvitamin D3 isomers were ca. ten times less potent than the natural hormone 1alpha,25-(OH)2D3 both in intestinal calcium transport and bone calcium mobilization.  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