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 328 毫秒
1.
Basic fibroblast growth factor (bFGF) binds to cell surface receptors and to heparin sulfate proteoglycans. Heparan sulfate binding may limit bFGF degradation and be an obligatory step for bFGF cell interaction. Transforming growth factor-beta 1 (TGF-beta 1) is a potent regulator of proteoglycan production and composition. The possibility that TGF-beta 1 synergistically regulates bFGF activity by altering bFGF-proteoglycan interactions was investigated. TGF-beta 1 increased 125I-bFGF binding to the extracellular matrix (ECM) of Balb/c3T3 cells 2-4-fold by increasing the number of bFGF binding sites. Increased bFGF binding correlated with a 2-5-fold increase in the production of sulfated proteoglycans, including heparan sulfate proteoglycans. TGF-beta 1 selectively stimulated production of high molecular mass proteoglycans (190-300 kDa) in conditioned medium and stimulated all proteoglycans in ECM. 125I-bFGF bound to TGF-beta 1 induced proteoglycans immobilized onto cationic nylon filters. Furthermore, ECM isolated from TGF-beta 1-treated cells incorporated more mitogenically active bFGF than native ECM. The mitogenic potential of the ECM was significantly reduced by treatment with heparinase. These results suggest that the ability of TGF-beta 1 to stimulate binding of bFGF to ECM, increase ECM heparan sulfate proteoglycan, and potentiate the mitogenic activity of bFGF are linked. Thus one aspect of TGF-beta 1/bFGF synergy may involve modulation of the ECM.  相似文献   

2.
Cell-surface proteoglycans have been known to be involved in many functions including interactions with components of the extracellular microenvironment, and act as co-receptors which bind and modify the action of various growth factors and cytokines. The purpose of this study was to determine the regulation by growth factors and cytokines on cell-surface proteoglycan gene expression in cultured human periodontal ligament (PDL) cells. Subconfluent, quiescent PDL cells were treated with various concentrations of serum, bFGF, PDGF-BB, TGF-beta1, IL-beta1, and IFN-gamma. RT-PCR technique was used, complemented with Northern blot for syndecan-1, to examine the effects of these agents on the mRNA expression of five cell-surface proteoglycans (syndecan-1, syndecan-2, syndecan-4, glypican and betaglycan). Syndecan-1 mRNA levels increased in response to serum, bFGF and PDGF-BB, but decreased in response to TGF-beta1, IL-1beta and IFN-gamma. In contrast, syndecan-2 mRNA levels were upregulated by TGF-beta1 and IL-1beta stimulation, but remained unchanged with the other agents. Betaglycan gene expression decreased in response to serum, but was upregulated by TGF-beta1 and unchanged by the other stimulants. Additionally, syndecan-4 and glypican were not significantly altered in response to the regulator molecules studied, with the exception that glypican is decreased in response to IFN-gamma. These data demonstrate that the gene expression of the five cell-surface proteoglycans studied is differentially regulated in PDL cells lending support to the notion of distinct functions for these cell-surface proteoglycans.  相似文献   

3.
Cultured bovine capillary endothelial (BCE) cells were found to synthesize and secrete high molecular mass heparan sulfate proteoglycans and glycosaminoglycans, which bound basic fibroblast growth factor (bFGF). The secreted heparan sulfate molecules were purified by DEAE cellulose chromatography, followed by Sepharose 4B chromatography and affinity chromatography on immobilized bFGF. Most of the heparinase-sensitive sulfated molecules secreted into the medium by BCE cells bound to immobilized bFGF at low salt concentrations. However, elution from bFGF with increasing salt concentrations demonstrated varying affinities for bFGF among the secreted heparan sulfate molecules, with part of the heparan sulfate requiring NaCl concentrations between 1.0 and 1.5 M for elution. Cell extracts prepared from BCE cells also contained a bFGF-binding heparan sulfate proteoglycan, which could be released from the intact cells by a short proteinase treatment. The purified bFGF-binding heparan sulfate competed with 125I-bFGF for binding to low-affinity binding sites but not to high-affinity sites on the cells. Heparan sulfate did not interfere with bFGF stimulation of plasminogen activator activity in BCE cells in agreement with its lack of effect on binding of 125I-bFGF to high-affinity sites. Soluble bFGF was readily degraded by plasmin, whereas bFGF bound to heparan sulfate was protected from proteolytic degradation. Treatment of the heparan sulfate with heparinase before addition of plasmin abolished the protection and resulted in degradation of bFGF by the added proteinase. The results suggest that heparan sulfate released either directly by cells or through proteolytic degradation of their extracellular milieu may act as carrier for bFGF and facilitate the diffusion of locally produced growth factor by competing with its binding to surrounding matrix structures. Simultaneously, the secreted heparan sulfate glycosaminoglycans protect the growth factor from proteolytic degradation by extracellular proteinases, which are abundant at sites of neovascularization or cell invasion.  相似文献   

4.
The effect of heparin on the rate of binding of basic fibroblast growth factor (bFGF) to high affinity (receptor) and low affinity (heparan sulfate) binding sites on endothelial cells and CHO cells transfected with FGF receptor-1 or FGF receptor-2 was investigated. Radiolabeled bFGF bound rapidly to both high and low affinity sites on all three types of cells. Addition of 10 micrograms/ml heparin eliminated binding to low affinity sites and decreased the rate of binding to high affinity sites to about 30% of the rate observed in the absence of heparin. However, the same amount of 125I-bFGF bound to high affinity sites at equilibrium in the presence and absence of heparin. The effect of heparin on the initial rate of binding to high affinity sites was related to the log of the heparin concentration. Depletion of the cells of heparan sulfates by treatment with heparinase also decreased the initial rate of binding to high affinity receptors. These results suggest that cell-surface heparan sulfates facilitate the interaction of bFGF with its receptor by concentrating bFGF at the cell surface. Dissociation rates for receptor-bound and heparan sulfate-bound bFGF were also measured. Dissociation from low affinity sites was rapid, with a half-time of 6 min for endothelial cell heparan sulfates and 0.5 min for Chinese hamster ovary heparan sulfates. In contrast, dissociation from receptors was slow, with a half-time of 46 min for endothelial cell receptors, 2.5 h for FGF receptor-1, and 1.4 h for FGF receptor-2. These results suggest that degradative enzymes may not be needed to release bFGF from the heparan sulfates in instances where receptors and heparan sulfate-bound bFGF are in close proximity because dissociation from heparan sulfates occurs rapidly enough to allow bFGF to bind to unoccupied receptors by laws of mass action.  相似文献   

5.
Heparan sulfate proteoglycans on the cell surface act as low affinity binding sites for acidic and basic fibroblast growth factor (FGF) [Moscatelli (1887): J Cell Physiol 131:123–130] and play an important role in the interaction of FGF with the FGF receptor (FGFR). In this study, several aspects of the interaction of FGFs with cell surface heparan sulfate proteoglycans were examined. Reciprocal cross blocking studies demonstrated that acidic FGF (aFGF) and basic FGF (bFGF) bind to identical or closely associated heparan sulfate motifs on BALB/c 3T3 cell surface heparan sulfate proteoglycans. However, the binding affinity of the two growth factros for these heparan sulfate proteoglycans differs considerably, competition binding data indicating that aFGF has a 4.7-fold lower affinity than bFGF for 3T3 heparan sulfate proteoglycan. Subsequent studies of dissociation kinetics demonstrated that bFGF dissociates form the FGFR at least 10-fold slower than aFGF, whereas, following removal of cell surface heparan sulfate proteoplycan. Subsequent studies of dissociation kinetic demonstrated that bFGF dissociates from the FGFR at least 10-fold slwer than aFGF, whereas, following removal of cell surface heparan sulfate proteoglycans by heparinase treatment, the dissociation rate of both FGFs is similar and rapid. These results support the concept that cell surface heparan sulfate proteoglycans stabilize the interactio fo FGF with FGFR, possibly by the formatin of a ternary complex. © Wiley-Liss, Inc.  相似文献   

6.
The role of cell density in modulating basic fibroblast growth factor binding and activity was investigated. A primary corneal stromal fibroblast cell culture system was used, since these cells do not constitutively express heparan sulfate proteoglycans in vivo except after injury. A 3-5-fold reduction in bFGF binding per cell was observed as cell density increased from 1000 to 35,000 cells/cm2. The cell density-dependent change in bFGF binding was not the result of altered FGFR expression as determined by equilibrium binding experiments and by immunoblot analysis. However, bFGF-cell surface receptor binding affinities were measured to be 10-20-fold higher at low cell densities than at intermediate and high cell density. bFGF-induced cell proliferation was also cell density-dependent, with maximal stimulation of proliferation 190-280% greater at intermediate densities (15,000 cells/cm2) than at other cell densities. This effect was specific to bFGF as serum, epidermal growth factor, and transforming growth factor-beta did not exhibit the same density-dependent profile. Further, heparan sulfate proteoglycans and, specifically, syndecan-4 were implicated as the modulator of bFGF binding and activity. Pretreatment of cell cultures with heparinase resulted in reduced bFGF binding to the cells and abrogated bFGF induced proliferation. These data suggest a mechanism by which cell density regulates heparan sulfate proteoglycan expression and modulates the cellular response to bFGF. Modulation of heparan sulfate proteoglycan expression might be an important aspect of the regulation of stromal cell migration and proliferation during wound healing.  相似文献   

7.
Numerous functions of heparan sulfate proteoglycans are mediated through interactions between their heparan sulfate glycosaminoglycan chains and extracellular ligands. Ligand binding specificity for some molecules, including many growth factors, is determined by complex heparan sulfate fine structure, where highly sulfated, iduronate-rich domains alternate with N-acetylated domains. Syndecan-4, a cell surface heparan sulfate proteoglycan, has a distinct role in cell adhesion, suggesting its chains may differ from those of other cell surface proteoglycans. To determine whether the specific role of syndecan-4 correlates with a distinct heparan sulfate structure, we have analyzed heparan sulfate chains from the different surface proteoglycans of a single fibroblast strain and compared their ability to bind the Hep II domain of fibronectin, a ligand known to promote focal adhesion formation through syndecan-4. Despite distinct molecular masses of glypican and syndecan glycosaminoglycans and minor differences in disaccharide composition and sulfation pattern, the overall proportion and distribution of sulfated regions and the affinity for the Hep II domain were similar. Therefore, adhesion regulation requires core protein determinants of syndecan-4.  相似文献   

8.
Kininogens, the high molecular weight precursor of vasoactive kinins, bind to a wide variety of cells in a specific, reversible, and saturable manner. The cell docking sites have been mapped to domains D3 and D5(H) of kininogens; however, the corresponding cellular acceptor sites are not fully established. To characterize the major cell binding sites for kininogens exposed by the endothelial cell line EA.hy926, we digested intact cells with trypsin and other proteases and found a time- and concentration-dependent loss of (125)I-labeled high molecular weight kininogen (H-kininogen) binding capacity (up to 82%), indicating that proteins are crucially involved in kininogen cell attachment. Cell surface digestion with heparinases similarly reduced kininogen binding capacity (up to 78%), and the combined action of heparinases and trypsin almost eliminated kininogen binding (up to 85%), suggesting that proteoglycans of the heparan sulfate type are intimately involved. Consistently, inhibitors such as p-nitrophenyl-beta-d-xylopyranoside and chlorate interfering with heparan sulfate proteoglycan biosynthesis reduced the total number of kininogen binding sites in a time- and concentration-dependent manner (up to 67%). In vitro binding studies demonstrated that biotinylated H-kininogen binds to heparan sulfate glycosaminoglycans via domains D3 and D5(H) and that the presence of Zn(2+) promotes this association. Cloning and over-expression of the major endothelial heparan sulfate-type proteoglycans syndecan-1, syndecan-2, syndecan-4, and glypican in HEK293t cells significantly increased total heparan sulfate at the cell surface and thus the number of kininogen binding sites (up to 3. 3-fold). This gain in kininogen binding capacity was completely abolished by treating transfected cells with heparinases. We conclude that heparan sulfate proteoglycans on the surface of endothelial cells provide a platform for the local accumulation of kininogens on the vascular lining. This accumulation may allow the circumscribed release of short-lived kinins from their precursor molecules in close proximity to their sites of action.  相似文献   

9.
We have investigated the expression patterns and subcellular localization in nervous tissue of glypican, a major glycosylphosphatidylinositol-anchored heparan sulfate proteoglycan that is predominantly synthesized by neurons, and of biglycan, a small, leucine-rich chondroitin sulfate proteoglycan. By laser scanning confocal microscopy of rat central nervous tissue and C6 glioma cells, we found that a significant portion of the glypican and biglycan immunoreactivity colocalized with nuclear staining by propidium iodide and was also seen in isolated nuclei. In certain regions, staining was selective, insofar as glypican and biglycan immunoreactivity in the nucleus was seen predominantly in a subpopulation of large spinal cord neurons. The amino acid sequences of both proteoglycans contain potential nuclear localization signals, and these were demonstrated to be functional based on their ability to target β-galactosidase fusion proteins to the nuclei of transfected 293 cells. Nuclear localization of glypican β-galactosidase or Fc fusion proteins in transfected 293 cells and C6 glioma cells was greatly reduced or abolished after mutation of the basic amino acids or deletion of the sequence containing the nuclear localization signal, and no nuclear staining was seen in the case of heparan sulfate and chondroitin sulfate proteoglycans that do not possess a nuclear localization signal, such as syndecan-3 or decorin (which is closely related in structure to biglycan). Transfection of COS-1 cells with an epitope-tagged glypican cDNA demonstrated transport of the full-length proteoglycan to the nucleus, and there are also dynamic changes in the pattern of glypican immunoreactivity in the nucleus of C6 cells both during cell division and correlated with different phases of the cell cycle. Our data therefore suggest that in certain cells and central nervous system regions, glypican and biglycan may be involved in the regulation of cell division and survival by directly participating in nuclear processes.  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.
Human aortic endothelial cells (HAEC) and human umbilical vein endothelial cells (HUVEC) were labeled with 35SO(4)2- for 48 h. The membrane-associated proteoglycans were solubilized from these monolayers with detergent and purified by ion-exchange chromatography on Mono Q, incorporation in liposomes, and gel filtration. The liposome-intercalated proteoglycans were 125I-iodinated and treated with heparitinase before SDS-polyacrylamide gel electrophoresis. Radio-labeled proteins with apparent molecular masses of 130, 60, 46, 35, and 30 kDa (HAEC) and 180, 130, 62, 43, and 35 kDa (HUVEC) were detected by autoradiography. Further characterization by affinity chromatography on immobilized monoclonal antibodies and by Northern blot analysis provided evidence for the expression of syndecan, glypican, and fibroglycan in human endothelial cells. Most of the heparan sulfate which accumulated in the subendothelial matrix was implanted on a 400-kDa core protein. This protein was immunologically related to perlecan and bound to fibronectin. Binding studies on immobilized antithrombin III suggested that all membrane-associated heparan sulfate proteoglycan forms had the capacity to bind to antithrombin III but that high affinity binding was more typical for glypican. Most of the proteoglycans isolated from the extracellular matrix also bound only with low affinity to antithrombin III. These results imply that glypican may specifically contribute to the antithrombotic properties of the vascular wall.  相似文献   

12.
Fibroblast growth factor 2 (FGF2)-initiated FGF receptor (FGFR)-signaling requires the assistance of heparin/heparan sulfate. Here, we evaluated the effects of different heparan sulfate proteoglycan (HSPG)-expressing cell lines and HSPGs derived from these cells on FGF2-induced FGFR1-phosphorylation in heparan sulfate-negative BaF3 cells. HSPGs supplied in membrane-associated form, by presenting cells, were all effective promotors of FGF2-initiated FGFR1 phosphorylation, independently of their nature (syndecan/glypican) or cellular origin (human lung fibroblasts, transfected Namalwa cells, or transfected K562 cells). A treatment with heparitinase initially stimulated, but finally completely inhibited, the activity of these presenting cells. In comparison, equivalent amounts of soluble HSPGs, obtained by trypsinization of these cells or by immunopurification from cell extracts, did not promote FGF2-induced FGFR1-phosphorylation, yet removal of the less anionic species or a further treatment with heparitinase converted these soluble fractions into potent activators of FGF2/FGFR1 signaling. Extrapolating from current structural models, we suggest that FGFR dimerization and autophosphorylation is supported by cooperative "heparin-like end structures," and that cell surface association and concentration compensate for the relative scarcity of such end structures in native HSPGs. In this model, "proteolytic" shedding of heparan sulfate would act as a diluting, down-regulatory mechanism, while "heparanolytic" shedding might act as an up-regulatory mechanism, by increasing the concentration of these end structures.  相似文献   

13.
A receptor for acidic and basic fibroblast growth factors (aFGF and bFGF, respectively) was isolated from 7-day embryonic chick. Chromatography of solubilized membrane proteins on wheat germ agglutininagarose and aFGF-Sepharose yielded three major polypeptides migrating at 150, 70, and 45 kDa as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. These polypeptides were eluted from aFGF-Sepharose with either 1.0 M NaCl or 100 micrograms/ml heparin, but were not retained on underivatized Sepharose. Cross-linking of 125I-aFGF or 125I-bFGF to either crude membrane preparations or to purified fractions yielded a 165-kDa complex, suggesting the existence of a 150-kDa FGF receptor after subtraction of approximately 15 kDa for 125I-FGF. Addition of excess aFGF or bFGF competed for binding of either 125I-aFGF or 125I-bFGF to FGF receptor preparations. Purified FGF receptor fractions were subjected to sodium dodecyl sulfate-polyacrylamide gel electrophoresis, transferred to Immobilon membranes, and incubated with 125I-aFGF or 125I-bFGF in order to identify FGF-binding polypeptides. Bound 125I-aFGF and 125I-bFGF were displaced by aFGF and bFGF, but not epidermal growth factor, consistent with the identification of the 150-kDa polypeptide as a receptor for acidic and basic FGF. Treatment of purified FGF receptor fractions with N-glycanase demonstrated that the 150-kDa polypeptide contained approximately 10 kDa of N-linked oligosaccharide. The apparent molecular mass of the 150-kDa polypeptide was unaffected by treatment with heparitinase, indicating that the 150-kDa polypeptide is not a heparan sulfate proteoglycan. Together, these data suggest that the 150-kDa polypeptide is a FGF receptor that may mediate the biological activities of aFGF and bFGF.  相似文献   

14.
Treating the liposome-intercalatable heparan sulfate proteoglycans from human lung fibroblasts and mammary epithelial cells with heparitinase and chondroitinase ABC revealed different core protein patterns in the two cell types. Lung fibroblasts expressed heparan sulfate proteoglycans with core proteins of approximately 35, 48/90 (fibroglycan), 64 (glypican), and 125 kDa and traces of a hybrid proteoglycan which carried both heparan sulfate and chondroitin sulfate chains. The mammary epithelial cells, in contrast, expressed large amounts of a hybrid proteoglycan and heparan sulfate proteoglycans with core proteins of approximately 35 and 64 kDa, but the fibroglycan and 125-kDa cores were not detectable in these cells. Phosphatidylinositol-specific phospholipase C and monoclonal antibody (mAb) S1 identified the 64-kDa core proteins as glypican, whereas mAb 2E9, which also reacted with proteoglycan from mouse mammary epithelial cells, tentatively identified the hybrid proteoglycans as syndecan. The expression of syndecan in lung fibroblasts was confirmed by amplifying syndecan cDNA sequences from fibroblastic mRNA extracts and demonstrating the cross-reactivity of the encoded recombinant core protein with mAb 2E9. Northern blots failed to detect a message for fibroglycan in the mammary epithelial cells and in several other epithelial cell lines tested, while confirming the expression of both glypican and syndecan in these cells. Confluent fibroblasts expressed higher levels of syndecan mRNA than exponentially growing fibroblasts, but these levels remained lower than observed in epithelial cells. These data formally identify one of the cell surface proteoglycans of human lung fibroblasts as syndecan and indicate that the expression of the cell surface proteoglycans varies in different cell types and under different culture conditions.  相似文献   

15.
Bovine capillary endothelial (BCE) cells were incubated at 4 degrees C with 5 ng/ml 125I-basic fibroblast growth factor (bFGF) to equilibrate 125I-bFGF with high affinity cell surface receptors and low affinity matrix binding sites. 67% of the added 125I-bFGF bound to the matrix and 7% bound to receptors. The fate of bound bFGF was followed after cells were incubated in bFGF-free medium and were shifted to 37 degrees C to restore cell metabolism. 125I-bFGF bound to receptors decreased rapidly while the amount of 125I-bFGF bound to matrix was reduced more slowly. The rapid decrease in receptor-bound 125I-bFGF appeared to be due to a down-regulation of bFGF receptors; cells that had been treated for 5 h with bFGF had 60% fewer high affinity receptors than untreated cells. Despite the initial high level of 125I-bFGF binding to matrix, most of this 125I-bFGF was mobilized and metabolized by the cells. 125I-bFGF was internalized by the cells at 37 degrees C, leading to a constant accumulation of 125I-bFGF within the cell. Internalized bFGF was rapidly cleaved from an 18-kD form to a 16-kD form. The 16-kD form was more slowly degraded with a half-life of approximately 8 h. Degradation of internalized 125I-bFGF was inhibited by chloroquine, suggesting that the digestion occurred in a lysosomal compartment. The role of matrix binding sites in the internalization process was investigated. Binding to matrix sites seemed not to be directly involved in the internalization process, since addition of heparin at a concentration that blocked 95% of the binding to matrix had no effect on the initial rate of internalization of bFGF. BCE cells also released a substance that competed for the binding of bFGF to matrix but not to receptors. This substance bound to DEAE-cellulose and was sensitive to heparinase treatment, suggesting that it was a heparinlike molecule. Thus, heparinlike molecules produced by BCE cells can modulate the cellular interaction with bFGF. Matrix-associated heparinlike molecules bind bFGF which can later be metabolized by the cell, and secreted heparinlike molecules release bFGF from matrices.  相似文献   

16.
Several processes that occur in the luminal compartments of the tissues are modulated by heparin-like polysaccharides. To identify proteins responsible for the expression of heparan sulfate at the apex of polarized cells, we investigated the polarity of the expression of the cell surface heparan sulfate proteoglycans in CaCo-2 cells. Domain- specific biotinylation of the apical and basolateral membranes of these cells identified glypican, a GPI-linked heparan sulfate proteoglycan, as the major source of apical heparan sulfate. Yet, most of this proteoglycan was expressed at the basolateral surface, an unexpected finding for a glypiated protein. Metabolic labeling and chase experiments indicated that sorting mechanisms, rather than differential turnover, accounted for this bipolar expression of glypican. Chlorate treatment did not affect the polarity of the expression of glypican in CaCo-2 cells, and transfectant MDCK cells expressed wild-type glypican and a syndecan-4/glypican chimera also in an essentially unpolarized fashion. Yet, complete removal of the heparan sulfate glycanation sites from the glypican core protein resulted in the nearly exclusive apical targeting of glypican in the transfectants, whereas two- and one-chain mutant forms had intermediate distributions. These results indicate that glypican accounts for the expression of apical heparan sulfate, but that glycanation of the core protein antagonizes the activity of the apical sorting signal conveyed by the GPI anchor of this proteoglycan. A possible implication of these findings is that heparan sulfate glycanation may be a determinant of the subcellular expression of glypican. Alternatively, inverse glycanation-apical sorting relationships in glypican may insure near constant deliveries of HS to the apical compartment, or "active" GPI-mediated entry of heparan sulfate into apical membrane compartments may require the overriding of this antagonizing effect of the heparan sulfate chains.  相似文献   

17.
Fibroblast growth factor-2 (FGF2) is a potent angiogenic factor in gliomas. Heparan sulfate promotes ligand binding to receptor tyrosine kinase and regulates signaling. The goal of this study was to examine the contribution of heparan sulfate proteoglycans (HSPGs) to glioma angiogenesis. Here we show that all brain endothelial cell HSPGs carry heparan sulfate chains similarly capable of forming a ternary complex with FGF2 and fibroblast growth factor receptor-1c and of promoting a mitogenic signal. Immunohistochemical analysis revealed that glypican-1 was overexpressed in glioma vessel endothelial cells, whereas this cell-surface HSPG was consistently undetectable in normal brain vessels. To determine the effect of increased glypican-1 expression on FGF2 signaling, we transfected normal brain endothelial cells, which express low base-line levels of glypican-1, with this proteoglycan. Glypican-1 expression enhanced growth of brain endothelial cells and sensitized them to FGF2-induced mitogenesis despite the fact that glypican-1 remained a minor proteoglycan. In contrast, overexpression of syndecan-1 had no effect on growth or FGF2 sensitivity. We conclude that the glypican-1 core protein has a specific role in FGF2 signaling. Glypican-1 overexpression may contribute to angiogenesis and the radiation resistance characteristic of this malignancy.  相似文献   

18.
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   

19.
Fannon M  Forsten KE  Nugent MA 《Biochemistry》2000,39(6):1434-1445
Basic fibroblast growth factor (bFGF) binds to cell surface tyrosine kinase receptor proteins and to heparan sulfate proteoglycans. The interaction of bFGF with heparan sulfate on the cell surface has been demonstrated to impact receptor binding and biological activity. bFGF receptor binding affinity is reduced on cells that do not express heparan sulfate. The addition of soluble heparin or heparan sulfate has been demonstrated to rescue the bFGF receptor binding affinity on heparan sulfate deficient cells yet has also been shown to inhibit binding under some conditions. While the chemical requirements of the heparin-bFGF-receptor interactions have been studied in detail, the possibility that heparin enhances bFGF binding in part by physically associating with the cell surface has not been fully evaluated. In the study presented here, we have investigated the possibility that heparin binding to the cell surface might play a role in modulating bFGF receptor binding and activity. Balb/c3T3 cells were treated with various concentrations of sodium chlorate, so as to express a range of endogenous heparan sulfate sites, and [(125)I]bFGF binding was assessed in the presence of a range of heparin concentrations. Low concentrations of heparin (0.1-30 nM) enhanced bFGF receptor binding to an extent that was inversely proportional to the amount of endogenous heparan sulfate sites present. At high concentrations (10 microM), heparin inhibited bFGF receptor binding in cells under all conditions. The ability of heparin to stimulate and inhibit bFGF-receptor binding correlated with altered bFGF-stimulated tyrosine kinase activity and cell proliferation. Under control and chlorate-treated conditions, [(125) I]heparin was observed to bind with a high affinity to a large number of binding sites on the cells (K(d) = 57 and 50 nM with 3.5 x 10(6) and 3.6 x 10(6) sites/cell for control and chlorate-treated cells, respectively). A mathematical model of this process revealed that the dual functions of heparin in bFGF binding were accurately represented by heparin cell binding-mediated stimulation and soluble heparin-mediated inhibition of bFGF receptor binding.  相似文献   

20.
Heparin can activate a receptor tyrosine kinase.   总被引:3,自引:0,他引:3       下载免费PDF全文
G Gao  M Goldfarb 《The EMBO journal》1995,14(10):2183-2190
Heparin, a densely sulfated glycosaminoglycan produced by mast cells, is best known for its inhibitory effects on the blood coagulation system. Heparin or heparan sulfate proteoglycans are also essential cofactors for the interaction of fibroblast growth factors (FGFs) with their receptor tyrosine kinases (FGFRs). Here we show that heparin is a growth factor-independent activating ligand for FGFR-4. Heparin stimulates FGFR-4 autophosphorylation on transfected myoblasts, fibroblasts and lymphoid cells, and is most potent on cells lacking surface heparan proteoglycan. Two functional analogs of heparin, fucoidan and dextran sulfate, are also activators of FGFR-4, while neither heparin nor its analogs can stimulate FGFR-1 in the absence of FGF. A mutation in the FGFR-4 ectodomain which impairs receptor activation by FGFs does not interfere with activation by heparin, demonstrating that receptor domains required for heparin or FGF activation are not identical. Heparin activation of FGFR-4 or of a chimeric receptor bearing FGFR-4 ectodomain and FGFR-1 cytodomain triggers downstream tyrosine phosphorylation of several signaling proteins, and induces proliferation of cells bearing the chimeric receptor. Consistent with these findings, a soluble FGFR-4 ectodomain has strong FGF-independent affinity for immobilized heparin resin, while soluble FGFR-1 requires FGF for stable heparin interaction. Heparin activation of FGFR-4 is the first example of a mammalian polysaccharide serving as a signaling ligand.  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