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 890 毫秒
1.
The purpose of this study was to clarify the mechanism of the blood-brain barrier (BBB) transport of H-Tyr-D-Arg-Phe-beta-Ala-OH (TAPA), which is a novel dermorphin analog with high affinity for the micro 1-opioid receptor. The in vivo BBB permeation influx rate of [125I]TAPA after an i.v. bolus injection (7.3 pmol/g body weight) into mice was estimated to be 0.265 +/- 0.025 microL/(min.g of brain). The influx rate of [125I]TAPA was reduced 70% by the coadministration of unlabeled TAPA (33 nmol/g of brain), suggesting the existence of a specific transport system for TAPA at the BBB. In order to elucidate the BBB transport mechanism of TAPA, a conditionally immortalized mouse brain capillary endothelial cell line (TM-BBB4) was used as an in vitro model of the BBB. The acid-resistant binding of [125I]TAPA, which represents the internalization of the peptide into cells, was temperature- and concentration-dependent with a half-saturation constant of 10.0 +/- 1.7 microm. The acid-resistant binding of TAPA was significantly inhibited by 2,4-dinitrophenol, dansylcadaverine (an endocytosis inhibitor) and poly-l-lysine and protamine (polycations). These results suggest that TAPA is transported through the BBB by adsorptive-mediated endocytosis, which is triggered by binding of the peptide to negatively charged sites on the surface of brain capillary endothelial cells. Blood-brain barrier transport via adsorptive-mediated endocytosis plays a key role in the expression of the potent opioid activity of TAPA in the CNS.  相似文献   

2.
Heparan sulfate proteoglycans (HSPG) play a critical role in the formation of distinct fibroblast growth factor (FGF)-HS complexes, augmenting high-affinity binding and receptor activation. Perlecan, a secreted HSPG abundant in proliferating cells, is capable of inducing FGF-receptor interactions in vitro and angiogenesis in vivo. Stable and specific reduction of perlecan levels in mouse NIH 3T3 fibroblasts and human metastatic melanoma cells has been achieved by expression of antisense cDNA corresponding to the N-terminal and HS attachment domains of perlecan. Long-term perlecan downregulation is evidenced by reduced levels of perlecan mRNA and core protein as indicated by Northern blot analysis, immunoblots, and immunohistochemistry, using DNA probes and antibodies specific to mouse or human perlecan. The response of antisense perlecan-expressing cells to increasing concentrations of basic FGF (bFGF) is dramatically reduced in comparison to that in wild-type or vector-transfected cells, as measured by thymidine incorporation and rate of proliferation. Furthermore, receptor binding and affinity labeling of antisense perlecan-transfected cells with 125I-bFGF is markedly inhibited, indicating that eliminating perlecan expression results in reduced high-affinity bFGF binding. Both the binding and mitogenic response of antisense-perlecan-expressing clones to bFGF can be rescued by exogenous heparin or perlecan. These results support the notion that perlecan is a major accessory receptor for bFGF in mouse fibroblasts and human melanomas and point to the possible use of perlecan antisense constructs as specific modulators of bFGF-mediated responses.  相似文献   

3.
We have investigated the transport characteristics of dehydroepiandrosterone sulfate (DHEAS), a neuroactive steroid, at the blood-brain barrier (BBB) in a series of functional in vivo and in vitro studies. The apparent BBB efflux rate constant of [(3)H]DHEAS evaluated by the brain efflux index method was 2.68 x 10(-2) min(-1). DHEAS efflux transport was a saturable process with a Michaelis constant (K:(m)) of 32.6 microM: Significant amounts of [(3)H]DHEAS were determined in the jugular venous plasma by HPLC, providing direct evidence that most of the DHEAS is transported in intact form from brain to the circulating blood across the BBB. This efflux transport of [(3)H]DHEAS was significantly inhibited by common rat organic anion-transporting polypeptide (oatp) substrates such as taurocholate, cholate, sulfobromophthalein, and estrone-3-sulfate. Moreover, the apparent efflux clearance of [(3)H]DHEAS across the BBB (118 microl/min-g of brain) was 10.4-fold greater than its influx clearance estimated by the in situ brain perfusion technique (11.4 microl/min-g of brain), suggesting that DHEAS is predominantly transported from the brain to blood across the BBB. In cellular uptake studies using a conditionally immortalized mouse brain capillary endothelial cell line (TM-BBB4), [(3)H]DHEAS uptake by TM-BBB4 cells exhibited a concentration dependence with a K:(m) of 34.4 microM: and was significantly inhibited by the oatp2-specific substrate digoxin. Conversely, [(3)H]digoxin uptake by TM-BBB4 cells was significantly inhibited by DHEAS. Moreover, the net uptake of [(3)H]DHEAS at 30 min was significantly increased under ATP-depleted conditions, suggesting that an energy-dependent efflux process may also be involved in TM-BBB4. RT-PCR and sequence analysis suggest that an oatp2 is expressed in TM-BBB4 cells. In conclusion, DHEAS efflux transport takes place across the BBB, and studies involving in vitro DHEAS uptake and RT-PCR suggest that there is oatp2-mediated DHEAS transport at the BBB.  相似文献   

4.
5.
The binding of [125I]-epidermal growth factor (EGF) and [125I]-basic fibroblast growth factor (bFGF) to a number of single-cell cloned rat mammary cell lines was measured using a saturation assay. Similar numbers of high-affinity [125I]-EGF binding sites (KD 1.3 nM) were found in epithelial and myoepithelial-like cell lines. In contrast, high-affinity (KD 35-276 pM) [125I]-bFGF binding sites were present on fibroblastic and myoepithelial-like cell lines but were not detectable on epithelial cell lines. A series of cell lines representing stages in the differentiation pathway of epithelial cells to an elongated myoepithelial-like morphology showed a graded increase in the number of bFGF receptors. The sensitivity of a cell line to stimulation of DNA synthesis by bFGF correlated with the level of expression of bFGF receptors on the cellular surface. Complexes of cell surface receptors affinity-cross-linked to [125I]-bFGF were analysed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). In each case two distinct complexes having apparent molecular weights of 180 kDa and 160 kDa were observed.  相似文献   

6.
Human basement membrane heparan sulfate proteoglycan (HSPG) perlecan binds and activates fibroblast growth factor (FGF)-2 through its heparan sulfate (HS) chains. Here we show that perlecans immunopurified from three cellular sources possess different HS structures and subsequently different FGF-2 binding and activating capabilities. Perlecan isolated from human umbilical arterial endothelial cells (HUAEC) and a continuous endothelial cell line (C11 STH) bound similar amounts of FGF-2 either alone or complexed with FGFRalpha1-IIIc or FGFR3alpha-IIIc. Both perlecans stimulated the growth of BaF3 cell lines expressing FGFR1b/c; however, only HUAEC perlecan stimulated those cells expressing FGFR3c, suggesting that the source of perlecan confers FGF and FGFR binding specificity. Despite these differences in FGF-2 activation, the level of 2-O- and 6-O-sulfation was similar for both perlecans. Interestingly, perlecan isolated from a colon carcinoma cell line that was capable of binding FGF-2 was incapable of activating any BaF3 cell line unless the HS was removed from the protein core. The HS chains also exhibited greater bioactivity after digestion with heparinase III. Collectively, these data clearly demonstrate that the bioactivity of HS decorating a single PG is dependent on its cell source and that subtle changes in structure including secondary interactions have a profound effect on biological activity.  相似文献   

7.
The binding of [125I]-recombinant basic FGF (rec bFGF) to rat hepatic plasma membranes was investigated. [125I] rec bFGF bound to an apparent single class of high affinity binding sites (KD = 69 pM; Bmax = 9.61 fmoles/mg proteins). The absence of low affinity sites was confirmed by the inability of sulphated polysaccharides and heparinase to interfere with FGF binding. A good correlation existed between the ability of bovine pituitary-derived bFGF, rec bFGF and bovine brain-derived aFGF to displace [125I]rec bFGF from these binding sites and their in vitro potency on bovine aortic endothelial cell proliferation.  相似文献   

8.
A divalent cation-dependent association between heparin or heparan sulfate and the ectodomain of the FGF receptor kinase (FGFR) restricts FGF-independent trans-phosphorylation and supports the binding of activating FGF to self-associated FGFR. Here we show that in contrast to heparin, cellular heparan sulfate forms a binary complex with FGFR that discriminates between FGF-1 and FGF-2. FGFR type 4 (FGFR4) in liver parenchymal cells binds only FGF-1, whereas FGFR1 binds FGF-1 and FGF-2 equally. Cell-free complexes of heparin and recombinant FGFR4 bound FGF-1 and FGF-2 equally. However, in contrast to FGFR1, when recombinant FGFR4 was expressed back in epithelial cells by transfection, it failed to bind FGF-2 unless heparan sulfate was depressed by chlorate or heparinase treatment. Isolated heparan sulfate proteoglycan (HSPG) from liver cells in cell-free complexes with FGFR4 restored the specificity for FGF-1 and supported the binding of both FGF-1 and FGF-2 when complexed with FGFR1. In contrast, FGF-2 bound equally well to complexes of both FGFR1 and FGFR4 formed with endothelial cell-derived HSPG, but the endothelial HSPG was deficient for the binding of FGF-1 to both FGFR complexes. These data suggest that a heparan sulfate subunit is a cell type- and FGFR-specific determinant of the selectivity of the FGFR signaling complex for FGF. In a physiological context, the heparan sulfate subunit may limit the redundancy among the current 18 FGF polypeptides for the 4 known FGFR.  相似文献   

9.
Cultured bovine capillary endothelial (BCE) cells synthesize heparan sulfate proteoglycans (HSPG), which are both secreted into the culture medium and deposited in the cell layer. The nonsoluble HSPGs can be isolated as two predominant species: a larger 800-kD HSPG, which is recovered from preparations of extracellular matrix, and a 250-kD HSPG, which is solubilized by nonionic detergent extraction of the cells. Both HSPG species bind bFGF. 125I-bFGF bound to BCE cell cultures is readily released by either heparinase or plasmin. When released by plasmin, the growth factor is recovered from the incubation medium as a complex with the partly degraded high molecular mass HSPG. Endogenous bFGF activity is released by a proteolytic treatment of cultured BCE cells. The bFGF-binding HSPGs are also released when cultures are incubated with the inactive proenzyme plasminogen. Under such experimental conditions, the release of the extracellular proteoglycans can be enhanced by treating the cells either with bFGF, which increases the plasminogen activating activity expressed by the cells, or decreased by treating the cells with transforming growth factor beta, which decreases the plasminogen activating activity of the cells. Specific immune antibodies raised against bovine urokinase also block the release of HSPG from BCE cell cultures. We propose that this plasminogen activator-mediated proteolysis provides a mechanism for the release of biologically active bFGF-HSPG complexes from the extracellular matrix and that bFGF release can be regulated by the balance between factors affecting the pericellular proteolytic activity.  相似文献   

10.
We have investigated elastase-mediated alterations in the expression of basic fibroblast growth factor (bFGF) receptors and proteoglycan co-receptors and characterized the subsequent effects on bFGF receptor binding profiles. For these studies, pulmonary fibroblast cultures were treated with porcine pancreatic elastase, and elastase-mediated changes in bFGF receptor expression and binding profiles were assessed. Quantitation of [(35)S]sulfate-labeled proteoglycan and total glycosaminoglycan release from fibroblast matrices indicated that elastase treatment released sulfated proteoglycan from the cell surface in a time- and dose-dependent fashion that correlated strongly with elastase-mediated bFGF release. Ligand binding studies indicated that elastase treatment decreased total binding of (125)I-bFGF to the cell surface and affected both fibroblast growth factor receptor and heparan sulfate proteoglycan (HSPG) binding sites. Western blot analyses indicated that elastase treatment did not release significant amounts of fibroblast growth factor receptor protein. These findings indicate that elastase-mediated HSPG release from fibroblast matrices reduces the effective affinity of bFGF for its receptor. Collectively, these studies suggest that HSPG co-receptors are important mediators of the pulmonary fibroblast response to elastase treatment and that bFGF, HSPG, and other elastase-released entities play an important role in the response of the lung to chronic injury.  相似文献   

11.
Basic fibroblast growth factor (bFGF) was internalized at a rapid rate by Chinese hamster ovary (CHO) cells that do not express significant numbers of high affinity receptors for bFGF as well as CHO cells that have been transfected with cDNA encoding FGF receptor-1 or FGF receptor-2. Internalization of bFGF was completely blocked by the addition of 10 micrograms/ml heparin in the parental CHO cells but only partially inhibited in cells expressing transfected FGF receptors. Bovine aortic endothelial cells also exhibit heparin-sensitive and heparin-resistant internalization of bFGF. The internalization of bFGF through the heparin-resistant pathway in CHO cells was efficiently competed by addition of unlabeled bFGF, was proportional to the number of receptors expressed, and approached saturation, suggesting that the heparin-resistant internalization was due to high affinity receptors. Internalization of bFGF through the heparin-sensitive pathway was not efficiently competed by unlabeled bFGF and did not approach saturation at concentrations of bFGF up to 50 ng/ml, properties similar to the interaction of bFGF with low affinity heparan sulfate binding sites on the cell surface. Internalization of bFGF in CHO cells not expressing FGF receptors was inhibited by heparin, heparan sulfate, and dermatan sulfate, the same glycosaminoglycans that block binding to cell-surface heparin sulfates. Internalization of bFGF in the parental CHO cells was inhibited at the same concentrations of heparin that block binding to cell-surface heparan sulfates. Finally, inhibition of the sulfation of CHO cell heparan sulfates by the addition of chlorate or digestion of CHO cell heparan sulfates with heparinase inhibited bFGF internalization in the parental CHO cells. These results demonstrate that bFGF can be internalized through a direct interaction with cell-surface heparan sulfates. Thus, there are two pathways for internalization of bFGF: high affinity receptor-mediated and heparan sulfate-mediated.  相似文献   

12.
The formation of distinctive basic FGF-heparan sulfate complexes is essential for the binding of bFGF to its cognate receptor. In previous experiments, cell-surface heparan sulfate proteoglycans extracted from human lung fibroblasts could not be shown to promote high affinity binding of bFGF when added to heparan sulfate-deficient cells that express FGF receptor-1 (FGFR1) (Aviezer, D., D. Hecht, M. Safran, M. Eisinger, G. David, and A. Yayon. 1994. Cell 79:1005-1013). In alternative tests to establish whether cell-surface proteoglycans can support the formation of the required complexes, K562 cells were first transfected with the IIIc splice variant of FGFR1 and then transfected with constructs coding for either syndecan-1, syndecan-2, syndecan-4 or glypican, or with an antisense syndecan-4 construct. Cells cotransfected with receptor and proteoglycan showed a two- to three- fold increase in neutral salt-resistant specific 125I-bFGF binding in comparison to cells transfected with only receptor or cells cotransfected with receptor and anti-syndecan-4. Exogenous heparin enhanced the specific binding and affinity cross-linking of 125I-bFGF to FGFR1 in receptor transfectants that were not cotransfected with proteoglycan, but had no effect on this binding and decreased the yield of bFGFR cross-links in cells that were cotransfected with proteoglycan. Receptor-transfectant cells showed a decrease in glycophorin A expression when exposed to bFGF. This suppression was dose-dependent and obtained at significantly lower concentrations of bFGF in proteoglycan-cotransfected cells. Finally, complementary cell- free binding assays indicated that the affinity of 125I-bFGF for an immobilized FGFR1 ectodomain was increased threefold when the syndecan- 4 ectodomain was coimmobilized with receptor. Equimolar amounts of soluble syndecan-4 ectodomain, in contrast, had no effect on this binding. We conclude that, at least in K562 cells, syndecans and glypican can support bFGF-FGFR1 interactions and signaling, and that cell-surface association may augment their effectiveness.  相似文献   

13.
The clinical importance of somatostatin type-2 receptors (SSTR2) and the study of novel analogues of somatostatin such as OctreoScan or [Tyr3]-octreotide containing DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid) as metal chelator led us to develop a methodology to monitor the expression of SSTR2 on tumours of pancreatic origin (e.g. rat AR4-2J cancer cells). Usual binding assay protocols using the commercial [125I][Tyr1]-somatostatin radioligand failed, even in the presence of a cocktail of protease inhibitors with a broad spectrum of activity, possibly due to the high susceptibility of this tracer to proteases expressed in pancreatic cells. We prepared our own radioligand [125I][Tyr2]-octreotide which was shown to be much more resistant to degradation after incubation with AR4-2J plasma membranes. As expected, the increased stability of [125I][Tyr3]-octreotide was associated with good binding to SSTR2. Addition of appropriate protease inhibitors further increased the specific binding of [125I][Tyr3]-octreotide to AR4-2J plasma membranes without affecting the stability of the tracer, suggesting that the protease inhibitors also protect the integrity of SSTR2. Optimal conditions (time, temperature, medium) were developed for a binding assay in 96-well plates using AR4-2J plasma membranes in order to make the assay suitable for high-throughput analysis. This protocol was the basis for studying the in vivo regulation of SSTR2 expression in AR4-2J cells implanted into scid mice after exposure to different compounds.  相似文献   

14.
Lipoprotein lipase (LPL) is produced by cells in the artery wall and can mediate binding of lipoproteins to cell surface heparan sulfate proteoglycans (HSPG), resulting in endocytosis (the bridging function). Active, dimeric LPL may dissociate to inactive monomers, the main form found in plasma. We have studied binding/internalization of human low density lipoprotein (LDL), mediated by bovine LPL, using THP-1 monocytes and macrophages. Uptake of (125)I-LDL was similar in monocytes and macrophages and was not affected by the LDL-receptor family antagonist receptor-associated protein (RAP) or by the phagocytosis inhibitor cytochalasin D. In contrast, uptake depended on HSPG and on membrane cholesterol. Incubation in the presence of dexamethasone increased the endogenous production of LPL by the cells and also increased LPL-mediated binding of LDL to the cell surfaces. Monomeric LPL was bound to the cells mostly in a heparin-resistant fashion. We conclude that the uptake of LDL mediated by LPL dimers is receptor-independent and involves cholesterol-enriched membrane areas (lipid rafts). Dimeric and monomeric LPL differ in their ability to mediate binding/uptake of LDL, probably due to different mechanisms for binding/internalization.  相似文献   

15.
Heparan sulfate proteoglycans (HSPG) are ubiquitous constituents of mammalian cell surfaces and most extracellular matrices. A portion of the cell surface HSPG is anchored via a covalently linked glycosyl-phosphatidylinositol (Pl) residue, which can be released by treatment with a glycosyl-Pl specific phospholipase C (Pl-PLC). We report that exposure of bovine aortic endothelial and smooth muscle cells to Pl-PLC resulted in release of cell surface-associated, growth-promoting activity that was neutralized by antibasic fibroblast growth factor (bFGF) antibodies. Active bFGF was also released by treating the cells with bacterial heparitinase. Under the same conditions there was no release of mitogenic activity from cells (BHK-21, NIH/3T3, PF-HR9) that expressed little or no bFGF, as opposed to Pl-PLC-mediated release of active bFGF from the same cells transfected with the bFGF gene. The released bFGF competed with recombinant bFGF in a radioreceptor assay. Addition of Pl-PLC to sparsely seeded vascular endothelial cells resulted in a marked stimulation of cell proliferation, but there was no mitogenic effect of Pl-PLC on 3T3 fibroblasts. Studies with exogenously added 125I-bFGF revealed that about 6.5% and 20% of the cell surface-bound bFGF were released by treatment with Pl-PLC and heparitinase, respectively. Both enzymes also released sulfate-labeled heparan sulfate from metabolically labeled 3T3 fibroblasts. Pl-PLC failed to release 125I-bFGF from the subendothelial extracellular matrix (ECM), as compared to release of 60% of the ECM-bound bFGF by heparitinase. Our results indicate that 3-8% of the total cellular content of bFGF is associated with glycosyl-Pl anchored cell surface HSPG. This FGF may exert both autocrine and paracrine effects, provided that it is released by Pl-PLC and adequately presented to high affinity bFGF cell surface receptor sites.  相似文献   

16.
The distribution and nature of (somatostatin) SRIF receptors and receptor mRNAs was studied in the brain and periphery of various laboratory animals using in situ hybridisation, autoradiography and radioligand binding. The messenger RNA (mRNA) expression of SRIF receptors msst1, msst2, msst3, msst4 and msst5 was studied in the adult mouse brain by in situ hybridisation histochemistry using specific oligonucleotide probes and compared to that of adult rats. As observed in rat brain, sst3 receptor mRNA is prominently expressed across the mouse brain, although equivalent binding has not yet been identified in situ. Sst1 and sst2 receptor mRNA expression, was prominent and again comparable to that observed in rat brain, whereas sst4 and especially sst5 receptor mRNA show comparatively low levels, although the former appears to be widely distributed while the latter could only be identified in a few nuclei. Altogether, the data are compatible with current knowledge, i.e. sst1 and sst2 receptor mRNA is prominent (both receptors have been functionally identified in the brain and for sst2 in the periphery), sst3 mRNA is highly expressed but in the absence of any functional correlate remains elusive. The expression of sst4 mRNA is comparatively low (especially when compared to what is seen in the lung, where high densities of sst4 receptors are present) and it remains to be seen whether sst5 receptor mRNA, which is confined to a few nuclei, will play a role in the brain, keeping in mind that high levels are found in the pituitary. Radioligand binding studies were performed in CCL39 cells expressing the five human recombinant receptors and compared to binding in membranes of rat cerebral cortex with [125I]Tyr11-SRIF14 which in the presence of 120 mM labels primarily sst1 receptor as suggested by the better correlation hsst1 and similar rank order of potency. The profile of [125I]Tyr3-octreotide labelled sites in rat cortex correlates better with recombinant sst2 than sst3 or sst5 binding profiles. Finally, [125I]LTT-SRIF28-labelled sites in rat lung express a sst4 receptor profile in agreement with previous findings. SRIF receptor autoradiography was performed in the brain and peripheral tissue of rat and/or guinea-pig using a number of ligands known to label recombinant SRIF receptors: [125I]LTT-SRIF28, [125I]CGP 23996, [125I]Tyr10-CST, or [125I]Tyr3-octreotide. Although, [125I]Tyr10-CST has been shown to label all five recombinant SRIF receptors, it is apparent that this radioligand is not useful for autoradiographic studies. By contrast, the other three ligands show good signal to noise ratios in rat or guinea-pig brain, rat lung, rat pancreas, or guinea-pig ileum. In most tissues, [125I]Tyr3-octreotide represents a prominent part of the binding (when compared to [125I]LTT-SRIF28 and [125I]CGP 23996), suggesting that sst2 receptors are strongly expressed in most tissues; it is only in rat lung that [125I]LTT-SRIF28 and [125I]CGP 23996 show marked binding, whereas [125I]Tyr3-octreotide does apparently label no sites, in agreement with the sole presence of sst4 receptors in this tissue.  相似文献   

17.
We have studied the interaction of 125I-antithrombin (125I-AT) with microvascular endothelial cells (RFPEC) to localize the cellular site of anticoagulantly active heparan sulfate proteoglycans (HSPG). The radiolabeled protease inhibitor bound specifically to the above HSPG with a Kd of approximately 50 nM. Confluent monolayer RFPEC cultures exhibited a linear increase in the amount of AT bound per cell for up to 16 d, whereas suspension RFPEC cultures possessed a constant number of protease inhibitor binding sites per cell for up to 5 d. These results suggest that monolayer RFPEC cultures secrete anticoagulantly active HSPG, which then accumulate in the extracellular matrix. This hypothesis was confirmed by quantitative light and EM level autoradiography which demonstrated that the AT binding sites are predominantly located in the extracellular matrix with only small quantities of protease inhibitor complexed to the cell surface. We have also pinpointed the in vivo position of anticoagulantly active HSPG within the blood vessel wall. Rat aortas were perfused, in situ, with 125I-AT, and bound labeled protease inhibitor was localized by light and EM autoradiography. The anticoagulantly active HSPG were concentrated immediately beneath the aortic and vasa vasorum endothelium with only a very small extent of labeling noted on the luminal surface of the endothelial cells. Based upon the above data, we propose a model whereby luminal and abluminal anticoagulantly active HSPG regulate coagulation mechanism activity.  相似文献   

18.
In this study, GABA efflux transport from brain to blood was estimated by using the brain efflux index (BEI) method. [3H]GABA microinjected into parietal cortex area 2 (Par2) of the rat brain was eliminated from the brain with an apparent elimination half-life of 16.9 min. The blood-brain barrier (BBB) efflux clearance of [3H]GABA was at least 0.153 mL/min/g brain, which was calculated from the elimination rate constant (7.14 x 10(-2) x min(-1)) and the distribution volume in the brain (2.14 mL/g brain). Direct comparison of the apparent BBB influx clearance [3H]GABA (9.29 microL/min/g brain) and the apparent efflux clearance (153 microL/min/g brain) indicated that the efflux clearance was at least 16-fold greater than the influx clearance. In order to reduce the effect of metabolism in the neuronal cells following intracerebral microinjection, we determined the apparent efflux of [3H]GABA in the presence of nipecotic acid, a GABA transport inhibitor in parenchymal cells, using the BEI method. Under such conditions, the elimination of [3H]GABA across the BBB showed saturation and inhibition by probenecid in the presence of nipecotic acid. Furthermore, the uptake of [3H]GABA by MBEC4 cells was inhibited by GABA, taurine, beta-alanine and nipecotic acid in a concentration-dependent manner. It is likely that GABA inhibits the first step in the abluminal membrane uptake by brain endothelial cells, and that probenecid selectively inhibits the luminal membrane efflux transport process from the brain capillary endothelial cells based on the in vivo and in vitro evidence. The BBB acts as the efflux pump for GABA to reduce the brain interstitial fluid concentration.  相似文献   

19.
Promyelocytic leukemia HL-60 cells can be induced to differentiate to granulocytes, under the conditions of cultures in the presence of dimethyl sulfoxide (DMSO). Examination of the binding of 125I-labeled hemopexin to DMSO-induced HL-60 cells showed that the density of hemopexin receptors on the induced-cells was 1.35 times that on the uninduced cells. We proposed that a specific receptor for hemopexin was present on the plasma membranes of polymorphonuclear leukocytes (PMNs). The binding of human [125I]hemopexin to human PMNs at 4 degrees C was saturable with time and with increasing concentrations of [125I]hemopexin. Scatchard analysis of the binding revealed the presence of approximately 5.7 x 10(4) binding sites per cell with an apparent dissociation constant (Kd) of 2.3 x 10(-9) M. [125I]Hemopexin was rapidly bound then dissociated from the cells after the release of heme, when the cells were incubated with radioactive hemopexin at 37 degrees C. Incubation of the cells with the [59Fe]heme-hemopexin complex resulted in an accumulation of [59Fe]heme in the cells, with a temperature of 37 degrees C but not that of 4 degrees C. Ouabain or NaF inhibited not only the binding of [125I]hemopexin to PMNs but also the uptake of [59Fe]heme from [59Fe]heme hemopexin by the cells. Neither NH4 Cl nor chloroquine inhibited the uptake. Detergent extracts of 125I-labeled PMNs were incubated with a hemopexin-coupled Sepharose CL-6B. A polypeptide reacting with hemopexin-Sepharose was estimated to have a molecular weight of 80,000, as determined by polyacrylamide gel electrophoresis, in the presence of sodium dodecylsulfate. We propose that PMNs take up heme from hemopexin, as mediated by the 80,000 dalton receptor for hemopexin.  相似文献   

20.
A panel of four murine monoclonal IgG1 antibodies (mAbs) to a recombinant form of basic fibroblast growth factor (bFGF) was produced using somatic cell fusion techniques. Non-linear regression analysis of radioimmunoassay data for each mAb yielded the following dissociation constants (nM) for their interactions with bFGF: DE6 (0.822); AF11 (2.0); FE8 (2.31); and DG2 (20.0). One of the mAbs, DG2, was identified as a bFGF neutralizing antibody on the basis of its ability to inhibit, in vitro, the binding of [125I]-bFGF to high and low affinity bFGF sites on cultured baby hamster kidney cells and bFGF-induced [3H]-thymidine incorporation in cultured 3T3 cells, and in vivo, the angiogenic response to bFGF in a rat kidney capsule model of angiogenesis. The other mAbs displayed varying inhibitory activities in these assays. These mAbs, particularly DG2, may be well suited for a number of applications in bFGF research including immunoassays, immunohistochemical studies, and as functional antagonists of bFGF for examining its role in physiological processes such as reproduction, growth, and development.  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