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.
Previously we demonstrated that proliferation of peripheral blood mononuclear cells (PBMC) causes a five-fold increase in cellular uptake of biotin; this increase is mediated by an increased number of biotin transporters on the PBMC surface. In the present study, we investigated the specificity of this phenomenon by determining whether the cellular uptake of riboflavin also increases in proliferating PBMC and whether the increase is also mediated by an increased number of transporters per cell. We characterized [3H]riboflavin uptake in both quiescent and proliferating PBMC. In quiescent PBMC, [3H]riboflavin uptake exhibited saturation kinetics and was reduced by addition of unlabeled riboflavin (P < 0.05) or lumichrome (P < 0.01). These observations are consistent with transporter-mediated uptake. [3H]Riboflavin uptake was reduced at 4 degrees C compared with 37 degrees C (P < 0.01) and by 2, 4-dinitrophenol (P < 0.05) but not by ouabain or incubation in sodium-free medium. These data provide evidence for an energy-dependent but sodium-independent transporter. Proliferating PBMC accumulated approximately four times more [3H]riboflavin than quiescent PBMC (P < 0.05). Because both transporter affinity and transporter number per cell (as judged by maximal transport rate) were similar in quiescent and proliferating PBMC, we hypothesize that the increased riboflavin uptake by proliferating PBMC reflects only increased cellular volume. To test this hypothesis, PBMC volume was reduced using hyperosmolar medium; [3H]riboflavin uptake decreased to about 50% of isotonic controls (P < 0.01). Thus we conclude that proliferating PBMC increase cellular content of riboflavin and biotin by two different mechanisms.  相似文献   

2.
We sought todetermine whether the proliferation of immune cells affects thecellular uptake of the vitamin biotin. Peripheral blood mononuclearcells (PBMC) were isolated from healthy adults. The proliferationof PBMC was induced by either pokeweed lectin, concanavalin A, orphytohemagglutinin. When the medium contained a physiologicalconcentration of[3H]biotin,nonproliferating PBMC accumulated 406 ± 201 amol[3H]biotin · 106cells1 · 30 min1. For proliferatingPBMC, [3H]biotinuptake increased to between 330 and 722% of nonproliferating values.Maximal transport rates of[3H]biotin inproliferating PBMC were also about four times greater than those innonproliferating PBMC, suggesting that proliferation was associatedwith an increase in the number of biotin transporters on the PBMCmembrane. The biotin affinities and specificities of the transporterfor proliferating and nonproliferating PBMC were similar, providingevidence that the same transporter mediates biotin uptake in bothstates. [14C]ureauptake values for proliferating and nonproliferating PBMC were similar,suggesting that the increased[3H]biotin uptake wasnot caused by a global upregulation of transporters duringproliferation. We conclude that PBMC proliferation increases thecellular accumulation of biotin.  相似文献   

3.
It has been well established that human intestinal and liver epithelial cells transport biotin via an Na+-dependent carrier-mediated mechanism. The sodium-dependent multivitamin transport (SMVT), a biotin transporter, is expressed in both cell types. However, the relative contribution of SMVT toward total carrier-mediated uptake of physiological (nanomolar) concentrations of biotin by these cells is not clear. Addressing this issue is important, especially in light of the recent identification of a second human high-affinity biotin uptake mechanism that operates at the nanomolar range. Hence, we employed a physiological approach of characterizing biotin uptake by human-derived intestinal Caco-2 and HepG2 cells at the nanomolar concentration range. We also employed a molecular biology approach of selectively silencing the endogenous SMVT of these cells with specific small interfering RNAs (siRNAs), then examining carrier-mediated biotin uptake. The results showed that in both Caco-2 and HepG2 cells, the initial rate of biotin uptake as a function of concentration over the range of 0.1 to 50 nM to be linear. Furthermore, we found that the addition of 100 nM unlabeled biotin, desthiobiotin, or pantothenic acid to the incubation medium had no effect on the uptake of 2.6 nM [3H]biotin. Pretreatment of Caco-2 and HepG2 cells with SMVT specific siRNAs substantially reduced SMVT mRNA and protein levels. In addition, carrier-mediated [3H]biotin (2.6 nM) uptake by Caco-2 and HepG2 cells was severely (P 0.01) inhibited by the siRNAs pretreatment. These results demonstrate that the recently described human high-affinity biotin uptake system is not functional in intestinal and liver epithelial cells. In addition, the results provide strong evidence that SMVT is the major (if not the only) biotin uptake system that operates in these cells.  相似文献   

4.
A transporter present in intestinal cells and in choriocarcinoma cells has been shown to transport both pantothenic acid and biotin at similar transporter affinities. However, the concentration of pantothenic acid in most foods and biological fluids is approximately 200 times the concentration of biotin; theoretically, pantothenic acid might substantially reduce biotin transport via competition. In the present study, we sought to determine whether pantothenic acid reduces biotin transport by the biotin transporter in peripheral blood mononuclear cells (PBMC). PBMC were isolated from human blood by gradient centrifugation. Incubations with [(3)H]biotin and pantothenic acid were conducted at physiologic concentrations. Intracellular [(3)H]biotin was quantified after washing by liquid scintillation counting. Pantothenic acid at 10 to 1,000 nmol/L reduced biotin (475 pmol/L) uptake by less than 12% (P < 0.05). Based on Lineweaver-Burk plots, the competition was reversible. Several structural analogs of pantothenic acid at 1,000 nmol/L reduced biotin transport by only 7 to 15% (P = 0.13). No pattern of molecular structure required for recognition by the transporter was apparent. Extracellular pantothenic acid did not affect biotin efflux from [(3)H]biotin-loaded PBMC (P > 0.05), suggesting that countertransport of extracellular pantothenic acid and intracellular biotin does not increase biotin efflux from PBMC. We conclude that the physiologic effects of pantothenic acid on the transport of biotin in PBMC are likely to be quantitatively minor.  相似文献   

5.
Preparation of high-specific-activity D-[3-3H]pantothenic acid   总被引:1,自引:0,他引:1  
High-specific-activity D-[3-3H]pantothenic acid (5 Ci/mmol) was prepared from commercially available beta-[3-3H]alanine employing Escherichia coli strain DV1 (panD2 pan F1). This strain is defective in beta-alanine synthesis and pantothenate uptake, and under appropriate growth conditions converted 85 to 90% of the input beta-[3-3H]alanine to extracellular D-[3-3H]pantothenate. The radiolabeled vitamin was purified from the medium by thin-layer chromatography followed by reverse-phase high-performance liquid chromatography. The overall yield of D-[3-3H]pantothenic acid was 30% and radiochemical purity was greater than 99%.  相似文献   

6.
Development and Characterization of Pantothenic Acid Transport in Brain   总被引:1,自引:1,他引:0  
In vitro, the transport of [3H]pantothenic acid into and from rabbit brain slices was studied. In newborn rabbits and throughout development, forebrain and cerebellar slices were able to accumulate and phosphorylate [3H]pantothenic acid comparably to slices from adults. The accumulation and phosphorylation of [3H]pantothenic acid by adult forebrain slices were not decreased by substitution of LiCl for NaCl in the artificial CSF or by addition of short-chain fuels (e.g., 5 mM pyruvate or acetoacetate) to the medium. However, probenecid and ouabain (both 1 mM) and medium-chain fatty acids (e.g., 0.1 mM octanoate, nonanoate, and decanoate) profoundly inhibited [3H]pantothenic acid accumulation by forebrain slices but not intracellular phosphorylation and conversion to [3H]CoA. There in vitro results suggest that brain slices accumulate pantothenic acid by a saturable system (probably facilitated diffusion) that is sensitive to inhibition by probenecid and medium-chain fatty acids.  相似文献   

7.
The purpose of this study was to clarify the expression of Na+‐dependent multivitamin transporter (SLC5A6/SMVT) and its contribution to the supply of biotin and pantothenic acid to the human brain via the blood–brain barrier. DNA microarray and immunohistochemical analyses confirmed that SLC5A6 is expressed in microvessels of human brain. The absolute expression levels of SLC5A6 protein in isolated human and monkey brain microvessels were 1.19 and 0.597 fmol/μg protein, respectively, as determined by a quantitative targeted absolute proteomics technique. Using an antibody‐free method established by Kubo et al. (2015), we found that SLC5A6 was preferentially localized at the luminal membrane of brain capillary endothelium. Knock‐down analysis using SLC5A6 siRNA showed that SLC5A6 accounts for 88.7% and 98.6% of total [3H]biotin and [3H]pantothenic acid uptakes, respectively, by human cerebral microvascular endothelial cell line hCMEC/D3. SLC5A6‐mediated transport in hCMEC/D3 was markedly inhibited not only by biotin and pantothenic acid, but also by prostaglandin E2, lipoic acid, docosahexaenoic acid, indomethacin, ketoprofen, diclofenac, ibuprofen, phenylbutazone, and flurbiprofen. This study is the first to confirm expression of SLC5A6 in human brain microvessels and to provide evidence that SLC5A6 is a major contributor to luminal uptake of biotin and pantothenic acid at the human blood–brain barrier.

  相似文献   


8.
The uptake of natural lung surfactant into differentiated type II cells may be used for the targeted delivery of other molecules. The fluorescent anion pyranine [hydroxypyren-1,3,6-trisulfonic acid, sodium salt (HPTS)] was incorporated into a bovine surfactant labeled with [3H]dipalmitoylphosphatidylcholine ([3H]DPPC). The uptake of [3H]DPPC and of HPTS increased with time of incubation and concentration, decreased with the size of the vesicles used, and was stimulated by 8-bromo-cAMP and partially inhibited by hypertonic sucrose. However, the amount of HPTS uptake was approximately 100 times smaller than that of [3H]DPPC. This large difference was due to a more rapid regurgitation of some of the HPTS from the cells but not to leakage from the surfactant before uptake. The acidification of the internalized surfactant increased linearly over 90 min to 7.13, and after 24 h, a pH of 6.83 was measured. In conclusion, after internalization of a double-labeled natural surfactant, the lipid moieties were accumulated in relation to the anions, which were targeted to a compartment not very acidic and in part rapidly expelled from the cells.  相似文献   

9.
Developmental patterns and pharmacological and biochemical properties of taurine transport system were investigated using developing primary cultured neurons prepared from mouse cerebral cortex by trypsin treatment. [3H]Taurine was incorporated into neurons via a high-affinity transport system of which the Km value as well as the Vmax value increased during neuronal development in vitro. This transport system was also inhibited by sodium withdrawal from incubation medium and exposures for 15 h to several metabolic inhibitors such as 2,4-dinitrophenol and monoiodoacetate. In addition, [3H]taurine uptake in both neurons cultured for 3 and 14 days was competitively inhibited by beta-alanine, guanidinoethanesulfonate and hypotaurine. Cysteic acid and cysteine sulfinic acid, metabolic intermediates produced in the process of taurine biosynthesis in the brain from cysteine, induced significant reductions in [3H]taurine uptake in both types of cultured neurons, while cysteine, isethionic acid, cysteamine and cystamine exhibited no alterations in [3H]taurine transport. Moreover, non-competitive inhibition of [3H]taurine uptake by cysteic acid was observed in both neurons. These results clearly indicate that taurine uptake was mediated by the sodium- and energy-dependent transport system with high affinity in 14-day-old neurons as well as neurons cultured for 3 days and that both the Km and Vmax values of this transport system increase during neuronal development in vitro. The results described above suggest that the decrease in taurine content observed in developing brain is unlikely to be due to alteration in the capacity of the taurine transport system during neuronal development.  相似文献   

10.
Although liver fatty acid binding protein (L-FABP) is known to enhance uptake and esterification of straight-chain fatty acids such as palmitic acid and oleic acid, its effects on oxidation and further metabolism of branched-chain fatty acids such as phytanic acid are not completely understood. The present data demonstrate for the first time that expression of L-FABP enhanced initial rate and average maximal oxidation of [2,3-3H] phytanic acid 3.5- and 1.5-fold, respectively. This enhancement was not due to increased [2,3-3H] phytanic acid uptake, which was only slightly stimulated (20%) in L-FABP expressing cells after 30 min. Similarly, L-FABP also enhanced the average maximal oxidation of [9,10-3H] palmitic acid 2.2-fold after incubation for 30 min. However, the stimulation of L-FABP on palmitic acid oxidation nearly paralleled its 3.3-fold enhancement of uptake. To determine effects of metabolism on fatty acid uptake, a non-metabolizable fluorescent saturated fatty acid, BODIPY-C16, was examined by laser scanning confocal microscopy (LSCM). L-FABP expression enhanced uptake of BODIPY-C16 1.7-fold demonstrating that L-FABP enhanced saturated fatty acid uptake independent of metabolism. Finally, L-FABP expression did not significantly alter [2,3-3H] phytanic acid esterification, but increased [9,10-3H] palmitic acid esterification 4.5-fold, primarily into phospholipids (3.7-fold) and neutral lipids (9-fold). In summary, L-FABP expression enhanced branched-chain phytanic acid oxidation much more than either its uptake or esterification. These data demonstrate a potential role for L-FABP in the peroxisomal oxidation of branched-chain fatty acids in intact cells.  相似文献   

11.
Mouse ileal sodium dependent bile acid transporter (ISBT) was characterized using isolated enterocytes. Only enterocytes from the most distal portion showed Na+-dependent [3H]taurocholate uptake. Northern blot analysis using a probe against mouse ISBT revealed the expression of mouse ISBT mRNA to be restricted to the distal ileum. The Km and Vmax for Na+-dependent [3H]taurocholate transport into isolated ileocytes were calculated as 27 microM and 360 pmol/mg protein/min, respectively. Uptake of [3H]taurocholate was inhibited by N-ethylmaleimide. We have cloned ISBT cDNA from mouse ileum. The cDNA included the entire open reading frame coding 348 amino acid protein with seven hydrophobic segments and two N-glycosylation sites. COS-7 cells transfected with the expression vector containing this cDNA expressed Na+-dependent [3H]taurocholate uptake activity with a Km of 34 microM.  相似文献   

12.
In vitro, the transport of [14C]pantothenic acid into and from the isolated rabbit choroid plexus, an anatomical locus of the blood-CSF barrier, and brain slices was studied. The choroid plexus accumulated [14C]pantothenic acid from the medium against a concentration gradient, although at low concentrations (less than 1 microM) there was substantial intracellular phosphorylation and binding of the [14C]pantothenic acid. The saturable accumulation process in choroid plexus was inhibited by probenecid and caproic acid but not by nicotinic acid or by weak bases. The accumulation process was markedly inhibited by N-ethylmaleimide, poly-L-lysine (which blocks sodium transport), and low temperatures. [14C]Pantothenic acid was readily released from choroid plexus by a temperature-dependent process. Brain slices also accumulated and, at low concentrations, phosphorylated [14C]pantothenic acid from the medium by a temperature-, probenecid-, and N-ethylmaleimide-sensitive saturable process. However, unlike choroid plexus, brain slices did not concentrate free pantothenic acid and [14C]pantothenic acid accumulation was not sensitive to poly-L-lysine. [14C]Pantothenic acid was readily released from brain slices by a temperature-sensitive process. These results are consistent with the view that [14C]pantothenic acid enters the isolated choroid plexus and brain slices by active transport and facilitated diffusion, respectively.  相似文献   

13.
The norepinephrine transporter (NET) plays a pivotal role in terminating noradrenergic signaling and conserving norepinephrine (NE) through the process of re-uptake. Recent evidence suggests a close association between NE release and regulation of NET function. The present study evaluated the relationship between release and uptake, and the cellular mechanisms that govern these processes. KCl stimulation of PC12 cells robustly increased [3H]NE uptake via the NET and simultaneously increased [3H]NE release. KCl-stimulated increases in uptake and release were dependent on Ca2+. Treatment of cells with phorbol-12-myristate-13-acetate (PMA) or okadaic acid decreased [3H]NE uptake but did not block KCl-stimulated increases in [3H]NE uptake. In contrast, PMA increased [3H]NE release and augmented KCl-stimulated release, while okadaic acid had no effects on release. Inhibition of Ca2+-activated signaling cascades with KN93 (a Ca2+ calmodulin-dependent kinase inhibitor), or ML7 and ML9 (myosin light chain kinase inhibitors), reduced [3H]NE uptake and blocked KCl-stimulated increases in uptake. In contrast, KN93, ML7 and ML9 had no effect on KCl-stimulated [3H]NE release. KCl-stimulated increases in [3H]NE uptake were independent of transporter trafficking to the plasma membrane. While increases in both NE release and uptake mediated by KCl stimulation require Ca2+, different intracellular mechanisms mediate these two events.  相似文献   

14.
We have previously shown that 3,5,3'-triiodo-L-thyronine (L-T3) stimulates cell growth and a 4- to 8-fold increase in growth hormone mRNA in GH1 cells. These effects appear to be mediated by a thyroid hormone nuclear receptor with an equilibrium dissociation constant for L-T3 of 0.2 nM and an abundance of about 10,000 receptors per cell nucleus. In this report, we show that L-T3 exerts a pleiotypic effect on GH1 cells to rapidly (within 2 h) stimulate [3H]uridine uptake to a maximal value of 2.5- to 3-fold after 24 h. This results from an increase in the number of functional uridine "transport sites" as shown by studies documenting an increase in the apparent Vmax with no change in the Km, 17 microM. Although the labeling of the cellular uridine pool and pools of all phosphorylated uridine derivatives was increased by L-T3, there was no change in the relative amounts of the individual pools in cells incubated with or without hormone. The intracellular concentration of [3H]uridine was estimated to be similar to that of the medium, suggesting that facilitated transport mediates [3H]uridine uptake. That this increase in [3H]uridine transport was nuclear receptor-mediated is supported by the excellent correspondence of the L-T3 dose-response curve for [3H]uridine uptake and that for L-T3 binding to receptor. Finally, inhibition of protein synthesis by cycloheximide and RNA synthesis by actinomycin D demonstrated that the L-T3 effect required continuing protein and RNA synthesis. These results are consistent with an effect of the L-T3-nuclear receptor complex to increase uridine uptake in GH1 cells by altering the expression of gene(s) essential for the transport process.  相似文献   

15.
Transport of L-[3H]carnitine and acetyl-L-[3H]carnitine at the blood-brain barrier (BBB) was examined by using in vivo and in vitro models. In vivo brain uptake of acetyl-L-[3H]carnitine, determined by a rat brain perfusion technique, was decreased in the presence of unlabeled acetyl-L-carnitine and in the absence of sodium ions. Similar transport properties for L-[3H]carnitine and/or acetyl-L-[3H]carnitine were observed in primary cultured brain capillary endothelial cells (BCECs) of rat, mouse, human, porcine and bovine, and immortalized rat BCECs, RBEC1. Uptakes of L-[3H]carnitine and acetyl-L-[3H]carnitine by RBEC1 were sodium ion-dependent, saturable with K(m) values of 33.1 +/- 11.4 microM and 31.3 +/- 11.6 microM, respectively, and inhibited by carnitine analogs. These transport properties are consistent with those of carnitine transport by OCTN2. OCTN2 was confirmed to be expressed in rat and human BCECs by an RT-PCR method. Furthermore, the uptake of acetyl-L-[3H]carnitine by the BCECs of juvenile visceral steatosis (jvs) mouse, in which OCTN2 is functionally defective owing to a genetical missense mutation of one amino acid residue, was reduced. The brain distributions of L-[3H]carnitine and acetyl-L-[3H]carnitine in jvs mice were slightly lower than those of wild-type mice at 4 h after intravenous administration. These results suggest that OCTN2 is involved in transport of L-carnitine and acetyl-L-carnitine from the circulating blood to the brain across the BBB.  相似文献   

16.
Transport of taurine has been demonstrated in capillary preparations from adult rat brains using [3H]taurine. Taurine transport is mediated by a saturable high-affinity system which is entirely dependent on sodium ions. The apparent maximal influx (Vmax) and half-saturation concentration (Km) corresponded to 1.06.10(-4) mumol/min per mg protein and 27.5 microM, respectively. Competition experiments in the presence of sodium ion showed that [3H]taurine uptake was strongly inhibited by 0.1 mM unlabeled structural analogues of taurine such as beta-alanine and hypotaurine as well as unlabeled taurine. gamma-Aminobutyric acid (GABA) (0.1 mM) inhibited the uptake of labeled taurine by 30%, whereas isethionic acid, L-methionine, L-2,4-diaminobutyric acid, glycine, L-cysteinesulfonic acid and cystamine did not exhibit any inhibitory effect. The results suggest that the Na+ gradient is the principal source of energy for taurine transport into isolated brain capillaries. This transport system may play an active role in the regulation of taurine concentration in the brain extracellular space.  相似文献   

17.
The conversion of stearic acid to oleic acid (delta 9-desaturase) was followed in mouse thymocytes stimulated by either concanavalin A or concanavalin A + interleukin-2 resulting in different rates of cell proliferation. To estimate the plasma membrane turnover of oleic acid as compared to that of a saturated fatty acid, double-label experiments ([14C]oleic acid, [3H]palmitic acid) were performed. Following an inhibition delta 9-desaturase was found to be activated from the fourth hour of stimulation. In the early period of cell activation this process proved to be independent of protein synthesis, whereas in the stage of proliferation it was dependent on it. Increased membrane fluidity in the first 30 min of activation is not likely due to enrichment of oleic acid. Cell proliferation and microsomal desaturation seem to be coupled and an increasing amount of oleic acid is at least one of the factors resulting in increased fluidity of the surface membrane of proliferating cells.  相似文献   

18.
The association of [3H] [D-Ala2, D-Leu5] enkephalin ([3H]DADLE]) with mouse neuroblastoma cells (N4TG1) was investigated. Under identical conditions the time course, dose response curve and temperature dependence for ligand uptake were similar to those for ligand-induced receptor loss (down regulation). Uptake of [3H]DADLE was inhibited by opiate ligands as well as by the metabolic inhibitors sodium azide and 2,4 dinitrophenol. Comparison of the effects of these inhibitors on receptor binding, ligand uptake and receptor loss indicated that these cells accumulate [3H]DADLE in excess of their surface receptor number. The data suggest that receptor recycling occurs and that ligand is internalized via receptor mediated endocytosis.  相似文献   

19.
Intracellular glutathione was increased by 80% after exposure of bovine pulmonary arterial endothelial cells to 80% O2 (hyperoxia) for 24 h. No change in glutathione occurred in cells exposed to hypoxia (3% O2) for a corresponding period of time. The rate of uptake of [3H]glutamic acid also increased by 35-55% after 24 h of exposure of cells to hyperoxia, whereas exposure to hypoxia had no effect on the [3H]glutamic acid uptake. The increase in glutamic acid uptake reflected a specific effect on amino acid transport systems rather than a change in cell membrane permeability. The major portion of the increased uptake was inhibited by the elimination of sodium and the addition of the competitive inhibitor, cystine, to the incubation medium. Thus increases in glutamic acid uptake parallel increases in cellular glutathione, and glutamic acid may be a regulating factor in the increase in glutathione after exposure to hyperoxia.  相似文献   

20.
Intestinal absorption of biotin is mediated via the sodium-dependent multivitamin transporter (SMVT). Studies from our laboratory and others have characterized different aspects of the human SMVT (hSMVT), but nothing is currently known about protein(s) that may interact with hSMVT and affect its physiology/biology. In this study, a PDZ-containing protein PDZD11 was identified as an interacting partner with hSMVT using yeast two-hybrid screen of a human intestinal cDNA library. The interaction between hSMVT and PDZD11 was confirmed by in vitro GST-pull-down assay and in vivo in a mammalian cell environment by a two-hybrid luciferase and coimmunoprecipitation assays. Furthermore, confocal imaging of live human intestinal epithelial HuTu-80 cells expressing hSMVT-GFP and DsRed-PDZD11 demonstrated colocalization of these two proteins. We also examined the functional consequence of the interaction between hSMVT and PDZD11 in HuTu-80 cells and observed significant induction in [(3)H]biotin uptake upon coexpression of hSMVT and PDZD11. In contrast, knocking down of PDZD11 with gene-specific small interfering RNA led to a significant decrease in biotin uptake; biotinylation assay showed this to be associated with a marked decrease in level of expression of hSMVT at the cell membrane. By truncation approach, we also demonstrated that the PDZ binding domain that is located in the COOH-terminal tail of hSMVT polypeptide is involved in the interaction with PDZD11. These results demonstrate for the first time that PDZD11 is an interacting partner with hSMVT in intestinal epithelial cells and that this interaction affects hSMVT function and cell biology.  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