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.
LNCaP is an androgen-sensitive human prostatic cancer cell line. The effect of androgen on these cells is characterized by a bell-shaped growth response and a dose-dependent induction of prostate-specific antigen (PSA) production. The present study was carried out to gain further insight into the effect of androgen on LNCaP. Cells were cultured in phenol red-free RPMI-1640 supplemented with 10% charcoal-stripped fetal bovine serum, with concentrations of dihydrotestosterone (DHT) ranging from 0-10(-7) M, in a 4-day culture system. A bell-shaped growth response was reproduced with a peak level of cell count at 10(-10) M DHT. PSA secretion from these cells did not increase significantly until the DHT level in the medium reached 10(-9) M. A progressive increase in PSA secretion was observed at higher DHT concentrations accompanied with a progressive decline in cellular proliferation. The results of immunocytochemical analysis of PSA localization indicated that the proportion of cells with positive staining for PSA also increased with increasing concentrations of DHT. Analysis of androgen receptors, as determined by both immunocytochemistry and Western blot analysis, showed a decline in nuclear androgen receptor at low concentrations of DHT and an increase in the amount of receptor protein at high concentrations. These results indicated that the androgen-induced bell-shaped growth response in LNCaP cells represented the manifestation of two different cellular events in dose-related manner: cellular proliferation at low DHT concentrations and increased production of PSA at high DHT concentrations.  相似文献   

2.
1,25-dihydroxyvitamin D3 [1,25-(OH)2D3] is known to modulate the development of bone and other mesenchymal cell types. Since osteoblasts and adipocytes are thought to arise in bone marrow from a common progenitor, this work examined the effects of 1,25-(OH)2D3 on adipocyte development, and in particular on the expression of lipoprotein lipase (LPL), which is an early marker for the differentiated adipocyte. 3T3-L1 preadipocytes were cultured in the presence of 1,25-(OH)2D3 (10(-9) to 10(-7) M) for up to 7 days. LPL activity was measured in the medium and cell extracts, and LPL messenger RNA levels were measured by Northern blotting. When compared to control cells, 10(-7) M 1,25-(OH)2D3 increased medium LPL activity by 2- to 3-fold and cellular LPL by 1.5-fold. Significant increases in medium and cellular LPL were observed at 10(-9) M and were maximal at 10(-7) M. Along with the increase in LPL activity, there was an increase in LPL messenger RNA by 2-fold at 5 days, and by 5-fold at 7 days. In addition to an increase in LPL, 1,25-(OH)2D3 increased expression of aP2, an adipocyte-specific marker associated with differentiation. After the addition of 1,25-(OH)2D3, there was a decrease in 3T3-L1 cell number, which is consistent with differentiation, and a decrease in vitamin D receptors. Finally, these cells developed a different morphology. 1,25-(OH)2D3-treated cells assumed a rounded appearance, although without detachment from the dish and without the degree of lipid accumulation usually associated with the addition of insulin, isbutylmethylxanthine, and dexamethasone. It is concluded that 1,25-(OH)2D3 induced LPL expression in 3T3-L1 cells through an induction of differentiation-dependent mechanism(s). These findings suggest an important role for 1,25-(OH)2D3 in normal adipocyte differentiation.  相似文献   

3.
4.
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   

5.
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   

6.
We examined expression of the 1,25-dihydroxyvitamin D3 [1,25-(OH)2 D3] receptors in chromaffin cells of the adrenal medulla and the effects of 1,25(OH)2 D3 on expression of the tyrosine hydroxylase (TH) gene. Accumulation of 1,25(OH)2 D3 in the nuclei of adrenal medullary cells, but not in the adrenal cortex, was observed in mice intravenously injected with radioactively labeled hormone. 1,25(OH)2 D3 produced concentration-dependent increases in the TH mRNA levels in cultured bovine adrenal medullary cells (BAMC). The maximal increases (2-3-fold) occurred at 10(-8) M 1,25(OH)2 D3. Combined treatment with 1,25(OH)2 D3 and 20 microM nicotine had no additive effect on TH mRNA levels suggesting that transsynaptic (nicotinic) and vitamin D (hormonal) stimulation of TH gene expression are mediated through converging mechanisms. Induction of TH mRNA by 1,25(OH)2 D3 was not affected by calcium antagonist TMB-8. By increasing expression of the rate limiting enzyme in the catecholamine biosynthetic pathway, 1,25-(OH)2 D3 may participate in the regulation of catecholamine production in adrenal chromaffin cells. This regulation provides mechanisms through which 1,25(OH)2 D3 may control response and adaptation to stress.  相似文献   

7.
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   

8.
9.
Previous reports have suggested the involvement of voltage-activated calcium (Ca2+) channels in bone metabolism and in particular on the secretion of osteocalcin by osteoblast-like cells. We now report that potassium (K+) channels can also modulate the secretion of osteocalcin by MG-63 cells, a human osteosarcoma cell line. When 1,25-dihydroxyvitamin D3(1,25(OH)2D3)-treated MG-63 cells were depolarized by step increases of the extracellular K+ concentration ([K+]out) from 5-30 mM, osteocalcin (OC) secretion increased from a control value of 218 +/- 13 to 369 +/- 18 ng/mg of protein/48 h (p < 0.005 by analysis of variance). In contrast, in the absence of 1,25(OH)2D3, there is no osteocalcin secretion nor any effect of cell depolarization on this activity. The depolarization-induced increase in 1,25(OH)2D3-dependent osteocalcin secretion was totally inhibited in the presence of 10 microM Nitrendipine (a Ca2+ channel blocker, p < 0.005) without affecting cellular alkaline phosphatase nor cell growth. Charybdotoxin, a selective blocker of Ca2+-dependent K+ channels (maxi-K) present in MG-63 cells, stimulated 1,25(OH)2D3-induced osteocalcin synthesis about 2-fold (p < 0.005) after either 30, 60, or 120 minutes of treatment. However, Charybdotoxin was without effect on basal release of osteocalcin in the absence of 1,25(OH)2D3 pretreatment. Using patch clamp technique, we occasionally observed the presence of a small conductance K+ channel, compatible with an ATP-dependent K+ channel (GK[ATP]) in nonstimulated cells, whereas multiple channel openings were observed when cells were treated with Diazoxide, a sulfonamide derivative which opens GK(ATP). Western blot analysis revealed the presence of the N-terminal peptide of GK(ATP) in MG-63 cells, and its expression was regulated with the proliferation rate of these cells, maximal detection by Western blots being observed during the logarithmic phase of the cycle. Glipizide and Glybenclamide, selective sulfonylureas which can block GK(ATP), dose-dependently enhanced 1,25(OH)2D3-induced OC secretion (p < 0.005). Reducing the extracellular calcium concentration with EGTA (microM range) totally inhibited the effect of Glipizide and Glybenclamide on osteocalcin secretion (p < 0.005), which remained at the same levels as controls. Diazoxide totally prevented the effect of these sulfonylureas. These results suggest that voltage-activated Ca2+ channels triggered via cell depolarization can enhance 1,25(OH)2D3-induced OC release by MG-63 cells. In addition, OC secretion is increased by blocking two types of K+ channels: maxi-K channels, which normally hyperpolarize cells and close Ca2+ channels, and GK(ATP) channels. The role of these channels is closely linked to the extracellular Ca2+ concentration.  相似文献   

10.
Vascular endothelial growth factor (VEGF), a secreted endothelial cell-specific mitogen, is produced in endocrine organs and regulated by trophic hormones. Because angiogenesis and osteogenesis are closely regulated, we studied whether human osteoblast-like cells produce VEGF, and if so, what factors regulate VEGF mRNA expression. Human osteoblast-like cells (HObLC) derived from trabecular bone explants were cultured in alpha-MEM supplemented with 10% fetal calf serum. Northern blot analysis revealed that HObLC expressed VEGF mRNA, as did several human osteosarcoma cells. 1,25-(OH)2D3 increased the steady-state levels of VEGF mRNA in a time- and concentration-dependent manner in HObLC and one of the osteosarcoma cell lines, SaOS-2, accompanied by an increase in the concentration of immunoreactive VEGF in the conditioned medium. PTH and IGF-I also increased the level of VEGF mRNA in HObLC and SaOS-2 cells. Furthermore, 12-O-tetradecanoylphorbol ester stimulated VEGF mRNA in a time-and concentration-dependent manner. The VEGF mRNA expression induced by 1,25-(OH)2D3 was completely inhibited by H-7, but only partially by staurosporine. We have demonstrated that PTH, IGF-I, and most potently 1,25-(OH)2D3 stimulate the mRNA expression and secretion of VEGF in human osteoblast-like cells, suggesting that one of the anabolic effects of 1,25-(OH)2D3 on skeletal tissue may be mediated by VEGF produced by osteoblasts.  相似文献   

11.
Androgen plays a critical role in regulating the growth and differentiation of normal prostate epithelia, as well as the initial growth of prostate cancer cells. Nevertheless, prostate carcinomas eventually become androgen-unresponsive, and the cancer is refractory to hormonal therapy. To gain insight into the mechanism involved in this hormone-refractory phenomenon, we have examined the potential role of the androgen receptor (AR) in that process. We have investigated the expression of AR and two prostate-specific androgen-responsive antigens, prostatic acid phosphatase (PAcP) and prostate-specific antigen (PSA), for the functional activity of AR in LNCaP and PC-3 human prostate carcinoma cells. Our results are as follows. (i) Clone 33 LNCaP cells express AR, PAcP, and PSA, and cell growth is stimulated by 5alpha-dihydrotestosterone (DHT). Stimulation of cell growth correlates with decreased cellular PAcP activity. (ii) In clone 81 LNCaP cells, the expression of PAcP decreases with a concurrent decrease in the degree of androgen stimulation of cell growth, whereas the expression of PSA mRNA level is up-regulated by DHT, as in clone 33 cells. Conversely, in PAcP cDNA-transfected clone 81 cells, an additional expression of cellular PAcP correlates with an increased stimulation by androgen, higher than the corresponding control cells. (iii) PC-3 cells express a low level of functional AR with no detectable PAcP or PSA, and the growth of PC-3 cells is not affected by DHT treatment. Nevertheless, in two PAcP cDNA-transfected PC-3 sublines, the expression of exogenous cellular PAcP correlates with androgen stimulation. This androgen stimulation of cell growth concurs with an increased tyrosine phosphorylation of a phosphoprotein of 185 kDa. In summary, the data indicate that the expression of AR alone is not sufficient for androgen stimulation of cell growth. Furthermore, in AR-expressing prostate cancer cells, the expression of cellular PAcP correlates with androgen stimulation of cell proliferation.  相似文献   

12.
1,25-Dihydroxyvitamin D3 [1,25-(OH)2D3] inhibited DNA synthesis in transformed mouse keratinocytes (Pam212) in a time- and dose-dependent manner as measured by [3H]thymidine incorporation. To investigate the mechanism through which 1,25-(OH)2D3 acts, we examined its effects on Pam212 cells further transformed with the E1A oncogene. Here, we show that transformation of the cells with the E1A oncogene induced resistance to the effects of 1,25-(OH)2D3 on inhibition of growth of Pam212 cells. While 1,25-(OH)2D3 treatment increased the level of expression of vitamin D receptor mRNA 20-fold in parental cells, the E1A-transformed cells failed to express vitamin D receptor mRNA even after treatment with 1,25-(OH)2D3. Transfection of the E1A-transformed cell line with an expression construct encoding the vitamin D receptor restored receptor expression as well as the inhibition of growth by 1,25-(OH)2D3. These results suggest that one of the mechanisms for acquisition of 1,25-(OH)2D3 resistance induced by E1A may involve loss of vitamin D receptor inducibility by 1,25-(OH)2D3.  相似文献   

13.
14.
We investigated the role of the active form of vitamin D, 1,25-dihydroxyvitamin D3[1,25(OH)2D3], in promoting tumor cell invasiveness through the extracellular matrix, and showed that 1,25(OH)2D3-induced reduction of laminin production by the cells was correlated with the inhibitory effect of the hormone on tumor cell invasiveness. 1,25(OH)2D3 significantly inhibited invasiveness through the matrix, type IV collagenolytic and migratory activity, but not cell attachment to the matrix in human fibrosarcoma HT1080 cells. The 1,25(OH)2D3-induced inhibition showed the same dose dependency and magnitude for invasiveness as for the effects on type IV collagenolysis and cell migration. 1,25(OH)2D3 inhibited laminin production from the cells in a dose-dependent manner. The inhibitory effects of 1,25(OH)2D3 on the invasiveness, type IV collagenolysis and cell migration appeared to parallel the hormone-induced reduction of laminin production. Antilaminin monoclonal antibody, blocking the activity of laminin in the culture medium, inhibited HT1080 cell invasiveness. In the presence of exogenous laminin, 1,25(OH)2D3-induced inhibition of invasion was not observed. These findings suggest that 1,25(OH)2D3 acts on HT1080 cells, inhibiting the expression of laminin from the cells, and that the reduced laminin expression leads to the inhibition in the type IV collagenolytic and migratory activity of the cells, and consequently, to the inhibition of invasiveness through the extracellular matrix.  相似文献   

15.
Glial cell line-derived neurotrophic factor (GDNF) has significant therapeutic potentials, in particular for neurodegenerative disorders. To determine factors that would enhance GDNF expression, we analysed the effect of 1,25-(OH)2 D3 in C6 glioma cells. Treatment of C6 cells with 10(-7) M, 1,25-(OH)2 D3 for 48 h elicited an 18.5-fold increase in the level of GDNF mRNA. In addition, our results indicate that 1,25-(OH)2 D3 is effective at concentrations as low as 10(-10) M and that retinoic acid has additive effects. These data indicate that 1,25-(OH)2 D3 is a potent inducer of GDNF expression and suggest that 1,25-(OH)2 D3 may contribute to the regulation of GDNF in vivo.  相似文献   

16.
1,25-dihydroxyvitamin D3 (1,25(OH)2D3), gamma interferon (IFN-gamma) and granulocyte-macrophage colony-stimulating factor (GM-CSF) can regulate monocyte maturation and activation. Using the human monocytoid cell line U937, we have shown that these agents increase surface tumor necrosis factor (TNF) expression without directly affecting TNF release. GM-CSF and IFN-gamma combined with 1,25(OH)2D3 increased cellular TNF secretion to levels not seen with these agents alone. Ability to express and secrete TNF in part depended on degree of monocytic maturation. The combination of 1,25(OH)2D3 and GM-CSF, however, facilitated lipopolysaccharide (LPS)-mediated release of surface TNF from U937 cells, an effect that was temporally independent of maximal maturation. 1,25(OH)2D3 plus IFN-gamma was less effective than 1,25(OH)2D3 plus GM-CSF at facilitating TNF secretion. We postulate that 1,25(OH)2D3 and GM-CSF are required together to prime a specific mechanism, probably a protease, which cleaves TNF from the surface of monocytic cells. This protease, once primed, can be activated by a secondary stimulus such as LPS.  相似文献   

17.
Human and murine osteocalcin genes demonstrate similar cell-specific expression patterns despite significant differences in gene locus organization and sequence variations in cis-acting regulatory elements. To investigate whether differences in these regulatory regions result in an altered response to 1,25-dihydroxyvitamin D3 [1,25-(OH)2D3] in vivo, we compared the response of the endogenous mouse osteocalcin gene to a bacterial reporter gene directed by flanking regions of the human osteocalcin gene in transgenic mice. Transgene expression colocalized with endogenous osteocalcin expression in serial sections, being detected in osteoblasts, osteocytes and hypertrophic chondrocytes. In calvarial cell culture lysates from transgenic and nontransgenic mice, the endogenous mouse osteocalcin gene did not respond to 1,25-(OH)2D3 treatment. Despite this, transgene activity was significantly increased in the same cells. Similarly, Northern blots of total cellular RNA and in situ hybridization studies of transgenic animals demonstrated a maximal increase in transgene expression at 6 h after 1,25-(OH)2D3 injection (23.6+/-3.6-fold) with a return to levels equivalent to uninjected animals by 24 h (1.2+/-0.1-fold). This increase in transgene expression was also observed at 6 h after 1,25-(OH)2D3 treatment in animals on a low calcium diet (25.2+/-7.7-fold) as well as in transgenic mice fed a vitamin D-deficient diet containing strontium chloride to block endogenous 1,25-(OH)2D3 production (7.5+/-0.9-fold). In contrast to the increased transgene expression levels, neither endogenous mouse osteocalcin mRNA levels nor serum osteocalcin levels were significantly altered after 1,25-(OH)2D3 injection in transgenic or nontransgenic mice, regardless of dietary manipulations, supporting evidence for different mechanisms regulating the response of human and mouse osteocalcin genes to 1,25-(OH)2D3. Although the cis- and trans-acting mechanisms directing cell-specific gene expression appear to be conserved in the mouse and human osteocalcin genes, responsiveness to 1,25-(OH)2D3 is not. The mouse osteocalcin genes do not respond to 1,25-(OH)2D3 treatment, but the human osteocalcin-directed transgene is markedly upregulated under the same conditions and in the same cells. The divergent responses of these homologous genes to 1,25-(OH)2D3 are therefore likely to be due to differences in mouse and human osteocalcin-regulatory sequences rather than to variation in the complement of trans-acting factors present in mouse osteoblastic cells. Increased understanding of these murine-human differences in osteocalcin regulation may shed light on the function of osteocalcin and its regulation by vitamin D in bone physiology.  相似文献   

18.
Prior studies have shown that 24,25-(OH)2D3 and 1,25-(OH)2D3 regulate protein kinase C (PKC) in costochondral chondrocytes in a cell maturation-dependent manner, with 1,25-(OH)2D3 affecting primarily growth zone (GC) cells and 24,25-(OH)2D3 affecting primarily resting zone (RC) cells. In addition, 1,25-(OH)2D3 has been shown to increase phospholipase A2 activity in GC, while 24,25-(OH)2D3 has been shown to decrease phospholipase A2 activity in RC. Stimulation of phospholipase A2 in GC caused an increase in PKC, whereas inhibition of phospholipase A2 activity in RC cultures increased both basal and 24,25-(OH)2D3-induced PKC activity, suggesting that phospholipase A2 may play a central role in mediating the effects of the vitamin D metabolites on PKC. To test this hypothesis, RC and GC cells were cultured in the presence and absence of phospholipase A2 inhibitors (quinacrine and oleyloxyethylphosphorylcholine [OEPC]), phospholipase A2 activators (melittin and mastoparan), or arachidonic acid alone or in the presence of the target cell-specific vitamin D metabolite. PKC specific activity in the cell layer was determined as a function of time. Phospholipase A2 inhibitors decreased both basal and 1,25-(OH)2D3-induced PKC activity in GC. When phospholipase A2 activity was activated by inclusion of melittin or mastoparan in the cultures, basal PKC activity in RC was reduced, while that in GC was increased. Similarly, melittin and mastoparan decreased 24,25-(OH)2D3-induced PKC activity in RC and increased 1,25-(OH)2D3-induced PKC activity in GC. For both cell types, the addition of arachidonic acid to the culture media produced an effect on PKC activity that was similar to that observed when phospholipase A2 activators were added to the cells. These results demonstrate that vitamin D metabolite-induced changes in phospholipase A2 activity are directly related to changes in PKC activity. Similarly, exogenous arachidonic acid affects PKC in a manner consistent with activation of phospholipase A2. These effects are cell maturation- and time-dependent and metabolite-specific.  相似文献   

19.
It has been well established that human mononuclear phagocytes have the capacity to produce 1,25-dihydroxy-vitamin D3 [1,25(OH)3D3] and express the vitamin D receptor (VDR). However, 1 alpha-hydroxylase activity and VDR receptor expression during differentiation of monocytes (MO) into mature macrophages (MAC) have not been previously examined. The in vitro maturation of blood MO can serve as a model for the in vivo transformation of immature blood MO into MAC. Here, when cultured in the presence of serum, MO undergo characteristic changes in morphology, antigenic phenotype, and functional activity consistent with their differentiation into MAC. We serially measured 1,25(OH)2D3 and 24,25-dihydroxyvitamin D3 [24,25(OH)2D3] synthesis, specific [3H]-1,25(OH)2D3 binding, and VDR mRNA levels during in vitro maturation of MO into MAC and correlated these functions with maturation-associated changes in the phenotype (MAX.1 and CD71) and secretory repertoire (interleukin-1 beta [IL-1 beta], neopterin) of the cells. MO showed only little conversion of 25-(OH)D3 into 1,25(OH)2D3 (1.4 +/- 0.4 pmol/10(6) cells/6 h, n = 5) that increased gradually during maturation into MAC at day 8 of culture (5.3 +/- 4.3 pmol/10(6) cells/6 h, n = 5). Interferon-gamma (IFN-gamma) increased baseline 1,25(OH)2D3-synthesis approximately twofold during all phases of differentiation. The time course of increased 1,25(OH)2D3-synthesis correlated with enhanced secretion of neopterin and expression of MAX.1 and CD71. The addition of exogenous 1,25(OH)2D3 did not influence constitutive 1,25(OH)2D3 synthesis, but IFN-gamma-stimulated production was suppressed to baseline levels. Exogenous 1,25(OH)2D3 also stimulated 24,25(OH)2D3 synthesis in freshly isolated MO (from 1.0 +/- 0.8 pmol/6 h to 5.6 +/- 0.9 pmol), whereas matured MAC showed no 24,25(OH)2D3 synthesis. Furthermore, we examined the expression of the VDR during the differentiation process. VDR mRNA and protein were constitutively expressed in MO, whereas VDR was downregulated in mature MAC on both the mRNA and protein levels. Homologous upregulation of VDR protein by 1,25(OH)2D3 occurred in MO and, to a lesser degree, in MAC. In contrast, VDR mRNA concentrations were not influenced by 1,25(OH)2D3. Taken together, our results show that MO into MAC differentiation in vitro is associated with (1) an enhanced capacity to synthesize 1,25(OH)2D3, (2) a loss of 24,25(OH)2D3-synthesizing activity, and (3) a decrease in the expression of VDR mRNA and protein. Because 1,25(OH)2D3 was shown to induce differentiation of MO into MAC, our data sugest an autoregulatory mechanism of MO/MAC generation by 1,25(OH)2D3.  相似文献   

20.
This study examines the influence of chronic retroviral infection of mice with a LPBM5 virus mixture on the paracrine system involving immune cells and 1,25-(OH)2D3 in the spleen. Plasma ionized calcium, 25-(OH)D and 1,25-(OH)2D of infected mice were unchanged. In contrast, the specific binding of 1,25-(OH)2D3 to spleen cytosol and the number of monocyte/macrophages expressing 1,25-(OH)2D3 receptors (VDR) were markedly increased. The retroviral infection also influenced the local production of 1,25-(OH)2D3 in the spleen. It did not alter this production in monocyte/macrophages but increased that in isolated T cells. Isolated B cells in control mice did not produce 1,25-(OH)2D3, but they increased the ability of isolated T cells to produce this metabolite during coculture incubations. Infection altered this cell interaction as 1,25-(OH)2D3 production in infected T cells decreased when these cells were cocultured with infected B cells. Thus, chronic retroviral infection alters both the local vitamin D metabolism and VDR expression by immune cells in mice. These findings suggest close local interactions between 1,25-(OH)2D3 and immune system activation during retroviral infection.  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