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

5.
6.
Prostate cancer (PCa) cells express vitamin D receptors (VDR) and 1,25-dihydroxyvitamin D(3) (1,25(OH)(2)D(3)) inhibits the growth of epithelial cells derived from normal, benign prostate hyperplasia, and PCa as well as established PCa cell lines. The growth inhibitory effects of 1,25(OH)(2)D(3) in cell cultures are modulated tissue by the presence and activities of the enzymes 25-hydroxyvitamin D(3) 24-hydroxylase which initiates the inactivation of 1,25(OH)(2)D(3) and 25-hydroxyvitamin D(3) 1alpha-hydroxylase which catalyses its synthesis. In LNCaP human PCa cells 1,25(OH)(2)D(3) exerts antiproliferative activity predominantly by cell cycle arrest through the induction of IGF binding protein-3 (IGFBP-3) expression which in turn increases the levels of the cell cycle inhibitor p21 leading to growth arrest. cDNA microarray analyses of primary prostatic epithelial and PCa cells reveal that 1,25(OH)(2)D(3) regulates many target genes expanding the possible mechanisms of its anticancer activity and raising new potential therapeutic targets. Some of these target genes are involved in growth regulation, protection from oxidative stress, and cell-cell and cell-matrix interactions. A small clinical trial has shown that 1,25(OH)(2)D(3) can slow the rate of prostate specific antigen (PSA) rise in PCa patients demonstrating proof of concept that 1,25(OH)(2)D(3) exhibits therapeutic activity in men with PCa. Further investigation of the role of calcitriol and its analogs for the therapy or chemoprevention of PCa is currently being pursued.  相似文献   

7.
Vasoconstrictors activate phospholipase C (PLC), which hydrolyzes phosphatidylinositol 4,5-bisphosphate (PIP(2)), leading to calcium mobilization, protein kinase C activation, and contraction. Our aim was to investigate whether PLC-delta(1), a PLC isoform implicated in alpha(1)-adrenoreceptor signaling and the pathogenesis of hypertension, is involved in noradrenaline (NA) or endothelin (ET-1)-induced PIP(2) hydrolysis and contraction. Rat mesenteric small arteries were studied. Contractility was measured by pressure myography, phospholipids or inositol phosphates were measured by radiolabeling with (33)Pi or myo-[(3)H]inositol, and caveolae/rafts were prepared by discontinuous sucrose density centrifugation. PLC-delta(1) was localized by immunoblot analysis and neutralized by delivery of PLC-delta(1) antibody. The PLC inhibitor U73122, but not the negative control U-73342, markedly inhibited NA and ET-1 contraction but had no effect on potassium or phorbol ester contraction, implicating PLC activity in receptor-mediated smooth muscle contraction. PLC-delta(1) was present in caveolae/rafts, and NA, but not ET-1, stimulated a rapid twofold increase in PLC-delta(1) levels in these domains. PLC-delta(1) is calcium dependent, and removal of extracellular calcium prevented its association with caveolae/rafts in response to NA, concomitantly reducing NA-induced [(33)P]PIP(2) hydrolysis and [(3)H]inositol phosphate formation but with no effect on ET-1-induced [(33)P]PIP(2) hydrolysis. Neutralization of PLC-delta(1) by PLC-delta(1) antibody prevented its caveolae/raft association and attenuated the sustained contractile response to NA compared with control antibodies. In contrast, ET-1-induced contraction was not affected by PLC-delta(1) antibody. These results indicate the novel and selective role of caveolae/raft localized PLC-delta(1) in NA-induced PIP(2) hydrolysis and sustained contraction in intact vascular tissue.  相似文献   

8.
The rapid, nongenomic effects of 1alpha,25-dihydroxyvitamin D3 (1alpha,25-(OH)2D3 have been related to a 1,25D3-membrane associated, rapid response steroid binding protein or 1,25D3-[MARRS]bp, with a molecular weight of 65 kDa, in several tissues and species. Currently, no information is available concerning the nongenomic responses to 1alpha,25-(OH)2D3 in dental tissues. In order to investigate the expression of 1,25D3-[MARRS]bp in dental cells, in the presence or absence of 1alpha,25-(OH)2D3, we have used rabbit polyclonal antibodies directed against the N-terminus of the 1,25D3-[MARRS]bp (Ab099) that recognizes the 1alpha,25-(OH)2D3 binding protein in chick intestinal basolateral membranes and a mouse odontoblast-like cell line (MO6-G3). Western blotting and flow cytometric analyses with Ab099 specifically detected 1,25D3-[MARRS]bp in MO6-G3 cells. Moreover, 1,25D3-[MARRS]bp was up-regulated, in vivo, in differentiated dental cells. Electron microscopic analysis confirmed the plasma membrane localization of this binding protein and also showed its intracellular presence. Incubation of MO6-G3 cells with different doses of 1alpha,25-(OH)2D3 for 36 h resulted in an inhibition of 1,25D3-[MARRS]bp expression with a maximal effect at 50 nM steroid. In addition, the culture media of MO6-G3 cells contains immunoreactive 1,25D3-[MARRS]bp. Immunogold positive membrane vesicle-like structures are present in the extracellular matrix of MO6-G3 cells. Altogether, these results indicate that the 1,25D3-[MARRS]bp expression in MO6-G3 cells is modulated by 1alpha,25-(OH)2D3. In conclusion, this 1alpha,25-(OH)2D3 binding protein could play an important role in the rapid, nongenomic responses to 1alpha,25-(OH)2D3 in dental cells.  相似文献   

9.
This study tested the hypothesis that 1,25-dihydroxyvitamin D(3) [1,25(OH)(2)D(3)] plays a role in human umbilical vein endothelial cells (HUVEC) cultures. HUVEC were incubated with 10 or 100 nM 1,25(OH)(2)D(3) for 24 h, in the absence or presence of 40 ng/ml tumor necrosis factor-alpha (TNF-alpha) or 2 ng/ml interleukin-1alpha (IL-1alpha). 1,25(OH)(2)D(3) did not affect HUVEC viability and proliferation, while TNF-alpha, alone or in combination with the hormone, significantly inhibited HUVEC viability. [(3)H]thymidine incorporation in HUVEC treated with TNF-alpha or IL-1alpha significantly decreased, in the absence or in the presence of the hormone, while the levels of vitamin D receptor markedly increased in the presence of 1,25(OH)(2)D(3) alone or associated with TNF-alpha or IL-1alpha, in comparison to the control. The noteworthy increase in protein levels of intercellular adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1) induced by TNF-alpha was significantly decreased after incubation of the cells with 1,25(OH)(2)D(3), this effect not being seen on E-selectin expression. Neither apoptosis nor nuclear translocation of NF-kappaB, induced in HUVEC by TNF-alpha was influenced by 1,25(OH)(2)D(3) treatment.  相似文献   

10.
New insights into the mechanisms of vitamin D action   总被引:17,自引:0,他引:17  
  相似文献   

11.
We recently reported (Larsson and Nemere [2003]: Endocrinology 144:1726) the effects of growth and maturation on 1,25(OH)2D3-membrane initiated effects in the intestine of male chickens. Here we extend our observations to studies on females with two stages of high calcium demand: growth (7-14 weeks) and egg laying (28-58 weeks). The rapid stimulatory effect of 130 pM 1,25(OH)2D3 on calcium transport was assessed as a physiological response in perfused duodena of 7-, 14-, 28-, and 58-week-old chickens, and determined to be 308%, 184%, 170%, and 153%, respectively, of corresponding controls after 40 min. Saturation analyses of [3H]1,25(OH)2D3 binding to nuclear vitamin D receptor (VDR) indicated an absence of cooperative binding, no changes in dissociation constant (Kd) with age, and an increase in maximum binding capacity (Bmax) between 7-week birds and older age groups. Analyses of saturable binding of [3H]1,25(OH)2D3 to the membrane associated rapid response steroid binding protein (1,25D3-MARRS bp) in basal lateral membranes (BLM), indicated cooperative binding, and an increase in both Bmax and Kd with age. No changes in the age-related expression of 1,25D3-MARRS bp were found, as judged by Western analyses, suggesting that a shift in ligand binding to lower affinity membrane components accounted for the increase in calculated Bmax. Basal levels of protein kinase C (PKC) activity decreased with age, as did hormone enhancement of activity. Basal levels of protein kinase A (PKA) activity remained constant with age, while the magnitude of hormone stimulation increased. Comparison of dose-response curves for ion transport and kinase activities in 7-week chicks suggested that PKC mediates phosphate transport while PKA mediates calcium transport. Thus, the age-related loss of calcium transport is most likely related to loss of PKC-mediated phosphate transport.  相似文献   

12.
The steroid hormone 1,25-dihydroxyvitamin D3 [1,25(OH)2D3] rapidly stimulates the uptake of phosphate in isolated chick intestinal cells, while the steroid 24,25-dihydroxyvitamin D3 [24,25(OH)2D3] inhibits the rapid stimulation by 1,25(OH)2D3. Earlier work in this laboratory has indicated that a cellular binding protein for 24,25(OH)2D3 is the enzyme catalase. Since binding resulted in decreased catalase activity and increased H2O2 production, studies were undertaken to determine if pro-oxidant conditions mimicked the inhibitory actions of 24,25(OH)2D3, and anti-oxidant conditions prevented the inhibitory actions of 24,25(OH)2D3. An antibody against the 24,25(OH)2D3 binding protein was found to neutralize the inhibitory effect of the steroid on 1,25(OH)2D3-mediated 32P uptake. Incubation of cells in the presence of 50 nM catalase was also found to alleviate inhibition. In another series of experiments, isolated intestinal epithelial cells were incubated as controls or with 1,25(OH)2D3, each in the presence of the catalase inhibitor 3-amino-1,2,4-triazole, or with 1,25(OH)2D3 alone. Cells exposed to hormone alone again showed an increased accumulation of 32P, while cells treated with catalase inhibitor and hormone had uptake levels that were indistinguishable from controls. We tested whether inactivation of protein kinase C (PKC), the signaling pathway for 32P uptake, occurred. Incubation of cells with phorbol-13-myristate (PMA) increased 32P uptake, while cells pretreated with 50 microM H2O2 prior to PMA did not exhibit increased uptake. Likewise, PMA significantly increased PKC activity while cells exposed to H2O2 prior to PMA did not. It is concluded that catalase has a central role in mediating rapid responses to steroid hormones.  相似文献   

13.
Nemere I  Campbell K 《Steroids》2000,65(8):451-457
The effect of vitamin D status on levels of the putative 1, 25(OH)(2)D(3) membrane receptor (pmVDR) was studied in chick intestine, kidney, and brain. Western analyses and assays for specific [(3)H]1,25(OH)(2)D(3) binding indicated that, in intestine, pmVDR levels were greatest in -D chicks relative to +1,25D and +D animals (P < 0.05). In kidney, protein levels and specific binding followed the order +D > +1,25D, -D. In brain, vitamin D status did not affect protein levels or specific binding levels. In tissue from normal chicks, both protein and specific binding followed the order of intestine > kidney > brain membranes. Intestinal cells were further evaluated for the effect of 1,25(OH)(2)D(3) on selected "rapid responses." Extrusion of (45)Ca in response to 130 pM 1, 25(OH)(2)D(3) in vitro was greater in cells from -D chicks than from +1,25D or normal birds. Analyses of signal transduction events revealed diminished hormone-induced intracellular calcium oscillations (as assessed by fura-2 fluorescence), and lack of steroid-enhanced protein kinase (PK) A activity in intestinal epithelial cells from -D chicks relative to +D chicks. PK C activation by 130 pM 1,25(OH)(2)D(3) was approximately twofold in cells from +D or -D chicks. The combined results indicate that vitamin D status differentially affects the pmVDR in intestine, kidney, and brain. In intestine, vitamin D deficiency differentially affects (45)Ca handling, intracellular calcium oscillations, PK A and PK C activities in response to 1,25(OH)(2)D(3).  相似文献   

14.
1alpha,25(OH)(2)D(3) regulates rat growth plate chondrocytes via nuclear vitamin D receptor (1,25-nVDR) and membrane VDR (1,25-mVDR) mechanisms. To assess the relationship between the receptors, we examined the membrane response to 1alpha,25(OH)(2)D(3) in costochondral cartilage cells from wild type VDR(+/+) and VDR(-/-) mice, the latter lacking the 1,25-nVDR and exhibiting type II rickets and alopecia. Methods were developed for isolation and culture of cells from the resting zone (RC) and growth zone (GC, prehypertrophic and upper hypertrophic zones) of the costochondral cartilages from wild type and homozygous knockout mice. 1alpha,25(OH)(2)D(3) had no effect on [(3)H]-thymidine incorporation in VDR(-/-) GC cells, but it increased [(3)H]-thymidine incorporation in VDR(+/+) cells. Proteoglycan production was increased in cultures of both VDR(-/-) and VDR(+/+) cells, based on [(35)S]-sulfate incorporation. These effects were partially blocked by chelerythrine, which is a specific inhibitor of protein kinase C (PKC), indicating that PKC-signaling was involved. 1alpha,25(OH)(2)D(3) caused a 10-fold increase in PKC specific activity in VDR(-/-), and VDR(+/+) GC cells as early as 1 min, supporting this hypothesis. In contrast, 1alpha,25(OH)(2)D(3) had no effect on PKC activity in RC cells isolated from VDR(-/-) or VDR(+/+) mice and neither 1beta,25(OH)(2)D(3) nor 24R,25(OH)(2)D(3) affected PKC in GC cells from these mice. Phospholipase C (PLC) activity was also increased within 1 min in GC chondrocyte cultures treated with 1alpha,25(OH)(2)D(3). As noted previously for rat growth plate chondrocytes, 1alpha,25(OH)(2)D(3) mediated its increases in PKC and PLC activities in the VDR(-/-) GC cells through activation of phospholipase A(2) (PLA(2)). These responses to 1alpha,25(OH)(2)D(3) were blocked by antibodies to 1,25-MARRS, which is a [(3)H]-1,25(OH)(2)D(3) binding protein identified in chick enterocytes. 24R,25(OH)(2)D(3) regulated PKC in VDR(-/-) and VDR(+/+) RC cells. Wild type RC cells responded to 24R,25(OH)(2)D(3) with an increase in PKC, whereas treatment of RC cells from mice lacking a functional 1,25-nVDR caused a time-dependent decrease in PKC between 6 and 9 min. 24R,25(OH)(2)D(3) dependent PKC was mediated by phospholipase D, but not by PLC, as noted previously for rat RC cells treated with 24R,25(OH)(2)D(3). These results provide definitive evidence that there are two distinct receptors to 1alpha,25(OH)(2)D(3). 1alpha,25(OH)(2)D(3)-dependent regulation of DNA synthesis in GC cells requires the 1,25-nVDR, although other physiological responses to the vitamin D metabolite, such as proteoglycan sulfation, involve regulation via the 1,25-mVDR.  相似文献   

15.
1alpha,25-dihydroxy vitamin D3 has a major role in the regulation of the bone metabolism as it promotes the expression of key bone-related proteins in osteoblastic cells. In recent years it has become increasingly evident that in addition to its well-established genomic actions, 1alpha,25-dihydroxy vitamin D3 induces non-genomic responses by acting through a specific plasma membrane-associated receptor. Results from several groups suggest that the classical nuclear 1alpha,25-dihydroxy vitamin D3 receptor (VDR) is also responsible for these non-genomic actions of 1alpha,25-dihydroxy vitamin D3. Here, we have used siRNA to suppress the expression of VDR in osteoblastic cells and assessed the role of VDR in the non-genomic response to 1alpha,25-dihydroxy vitamin D3. We report that expression of the classic VDR in osteoblasts is required to generate a rapid 1alpha,25-dihydroxy vitamin D3-mediated increase in the intracellular Ca(2+) concentration, a hallmark of the non-genomic actions of 1alpha,25-dihydroxy vitamin D3 in these cells.  相似文献   

16.
The mechanisms of endogenous cannabinoid biosynthesis are not completely understood. We hypothesized that anandamide could be recycled by the cell to form new endocannabinoid molecules and released into the extracellular space. We determined that new endocannabinoids derived from exogenous anandamide or arachidonic acid were synthesized and released from RBL-2H3 cells in response to ionomycin. Treatment of RBL-2H3 cells with nystatin and progesterone, agents that disrupt organization of lipid raft/caveolae, resulted in the attenuation of anandamide and 2-arachidonyl glycerol synthesis and/or release in response to stimulation with ionomycin suggesting a role for these membrane microdomains in endocannabinoid biosynthesis. Furthermore, anandamide synthesis may be independent of N-acyl phosphatidylethanolamine phospholipase D as expression of the enzyme was not detected in RBL-2H3 cells. We also established that extracellular calcium is necessary for endocannabinoid biosynthesis because release of intracellular calcium stores alone does not promote endocannabinoid biosynthesis. Next, we examined the role of calcium as a 'switch' to activate the synthesis of anandamide and simultaneously reduce uptake. Indeed, [(3)H] anandamide uptake was reduced in the presence of calcium. Our findings suggest a mechanism indicative of calcium-modulated activation of anandamide synthesis and simultaneous termination of uptake.  相似文献   

17.
18.
Demonstrating 1,25(OH)2D3-stimulated calcium uptake in isolated chick intestinal epithelial cells has been complicated by simultaneous enhancement of both uptake and efflux. We now report that in intestinal cells of adult birds, or those of young birds cultured for 72 h, 1,25(OH)2D3-stimulates 45Ca uptake to greater than 140% of corresponding controls within 3 min of addition. Such cells have lost hormone-stimulated protein kinase C (PKC) activity, believed to mediate calcium efflux. To further test this hypothesis, freshly isolated cells were preincubated with calphostin C, and calcium uptake monitored in the presence or absence of steroid. Only cells treated with the PKC inhibitor demonstrated a significant increase in 45Ca uptake in response to 1,25(OH)2D3, relative to corresponding controls. In addition, phorbol ester was shown to stimulate efflux, while forskolin stimulated uptake. To further investigate the mechanisms involved in calcium uptake, we assessed the role of TRPV6 and its activation by beta-glucuronidase. beta-Glucuronidase secretion from isolated intestinal epithelial cells was significantly increased by treatment with 1,25(OH)2D3, PTH, or forskolin, but not by phorbol ester. Treatment of cells with beta-glucuronidase, in turn, stimulated 45Ca uptake. Finally, transfection of cells with siRNA to either beta-glucuronidase or TRPV6 abolished 1,25(OH)2D3-enhanced calcium uptake relative to controls transfected with scrambled siRNA. Confocal microscopy further indicated rapid redistribution of enzyme and calcium channel after steroid. 1,25(OH)2D3 and PTH increase calcium uptake by stimulating the PKA pathway to release beta-glucuronidase, which in turn activates TRPV6. 1,25(OH)2D3-enhanced calcium efflux is mediated by the PKC pathway.  相似文献   

19.
Long-chain fatty acid uptake into adipocytes depends on lipid raft function   总被引:7,自引:0,他引:7  
This study investigates the role of lipid rafts and caveolae, a subclass of lipid raft microdomains, in the binding and uptake of long-chain fatty acids (LCFA) by 3T3-L1 cells during differentiation. Disruption of lipid rafts by beta-cyclodextrin (betaCD) or selective inhibition of caveolae by overexpression of a dominant-negative mutant of caveolin-3 (Cav(DGV)) resulted in disassembly of caveolae structures at the cell surface, as assessed by electron microscopy. While in 3T3-L1 fibroblasts, which express few caveolae, Cav(DGV) or betaCD had no effect on LCFA uptake, in 3T3-L1 adipocytes the same treatments decreased the level of [(3)H]oleic acid uptake by up to 55 +/- 8 and 49 +/- 7%, respectively. In contrast, cholesterol loading of 3T3-L1 adipocytes resulted in a 4-fold increase in the extent of caveolin-1 expression and a 1.7-fold increase in the level of LCFA uptake. Both the inhibitory and enhancing effects of these treatments were constantly increasing with the [(3)H]oleic acid incubation time up to 5 min. Incubation of 3T3-L1 adipocytes with [(3)H]stearate followed by isolation of a caveolin-1 positive detergent-resistant membrane (DRM) fraction revealed that [(3)H]stearate binds to caveolae. Fatty acid translocase (FAT/CD36) was found to be present in this DRM fraction as well. Our data thus strongly indicate a critical involvement of lipid rafts in the binding and uptake of LCFA into 3T3-L1 adipocytes. Furthermore, our findings suggest that caveolae play a pivotal role in lipid raft-dependent LCFA uptake. This transport mechanism is induced in conjunction with cell differentiation and might be mediated by FAT/CD36.  相似文献   

20.
The vitamin D metabolite, 24R,25-dihydroxyvitamin D(3) (24R,25(OH)(2)D(3)), was tested for its ability to specifically bind to basal lateral membranes isolated from intestinal epithelium of Atlantic cod (a seawater fish), carp (a freshwater fish), and chicken. Specific saturable binding was demonstrated in membranes from all three species. Membranes from Atlantic cod, carp, and chicken revealed K(d)'s of 7.3 +/- 0.9, 12.5 +/- 0.9 and 7.8 +/- 0.1 nM, and a B(max) for each species estimated to 57.9 +/- 2.9, 195.1 +/- 8.4 and 175 +/- 0.8 fmol/mg protein, respectively. Scatchard analyses indicated a convex curvature and Hill analyses revealed apparent Hill coefficients of 1.84 +/- 0.28, 1.80 +/- 0.29, and 1.78 +/- 0.27 for Atlantic cod, carp and chicken, suggesting a positive cooperative binding in all three species. Basal lateral membranes from Atlantic cod and carp were used to further characterize the binding moiety. In competition studies, basal lateral membranes from Atlantic cod or carp did not discriminate between 24R,25(OH)(2)D(3) and the 24S,25(OH)(2)D(3) isomer, whereas, 1,25(OH)(2)D(3) and 25(OH)D(3), were less effective in competing with [(3)H]24R,25(OH)(2)D(3) for binding to basal lateral membranes in Atlantic cod and carp. In both the Atlantic cod and carp enterocyte basal lateral membranes, the binding activity could be extracted equally well with high salt as with detergent, indicating a peripheral membrane protein rather than an integral membrane binding protein. Finally, isolated Atlantic cod and carp enterocytes were chosen for analyses of signal transduction events mediated by the putative receptor. In both species, 24R,25(OH)(2)D(3) but not 24S,25(OH)(2)D(3), suppressed Ca(2+)-uptake by enterocytes in a dose-dependent manner. Enterocytes from Atlantic cod and carp, acclimated to Ca(2+)-free media, responded by an intracellular Ca(2+)-release within seconds after addition of 24R,25(OH)(2)D(3) or 24S,25(OH)(2)D(3). The effects on intracellular Ca(2+)-release were dose-dependent for both metabolites. 24S,25(OH)(2)D(3) was effective at lower concentrations and triggered a higher response compared to 24R,25(OH)(2)D(3). These results suggest that the binding molecule(s) for 24R,25(OH)(2)D(3) and 24S,25(OH)(2)D(3) is/are capable of acting as a receptor, mediating rapid, non-genomic responses in intestinal cells.  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