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 15 毫秒
1.
Heparin and heparin-like molecules may function, apart from their effect on hemostasis, as regulators of cell growth and neovascularization. We investigated whether similar effects are exerted by laminarin sulfate, an unrelated polysulfated saccharide isolated from the cell wall of seaweed and composed of chemically O-sulfated b?-(1,3)-linked glucose residues. Laminarin sulfate exhibits about 30% of the anticoagulant activity of heparin and is effective therapeutically in the prevention and treatment of cerebrovascular diseases. We characterized the effect of laminarin sulfate on interaction of the heparin-binding angiogenic factor, basic fibroblast growth factor (bFGF), with a naturally produced subendothelial extracellular matrix (ECM) and with cell surface receptor sites. Laminarin sulfate (1-2 m?g/ml) inhibited the binding of bFGF to ECM and to the surface of vascular smooth muscle cells (SMC) in a manner similar to that observed with heparin. Likewise, laminarin sulfate efficiently displaced both ECM-and cell-bound bFGF at concentrations as low as 1 m?g/ml. Both laminarin sulfate and heparin efficiently induced restoration of bFGF receptor binding in xylosyltransferase-deficient CHO cell mutants defective in initiation of glycosaminoglycan synthesis. Moreover, laminarin sulfate elicited bFGF receptor activation and mitogenic response in heparan sulfate(HS)-deficient, cytokine-dependent lymphoid cells. These results indicate that laminarin sulfate effectively replaced the need for heparin and HS in the induction of bFGF receptor binding and signaling. In other experiments, laminarin sulfate was found to inhibit the proliferation of vascular SMC in a manner similar to that observed with heparin. These effects of laminarin sulfate may have potential clinical applications in diverse situations such as wound healing, angiogenesis, and atherosclerosis. © 1995 Wiley-Liss, Inc.  相似文献   

2.
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   

3.
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   

4.
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   

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.
A Yayon  M Klagsbrun  J D Esko  P Leder  D M Ornitz 《Cell》1991,64(4):841-848
The role of low affinity, heparin-like binding sites for basic fibroblast growth factor (bFGF) was investigated in CHO cells mutant in their metabolism of glycosaminoglycans. Heparan sulfate-deficient mutants transfected to express a cloned mouse FGF receptor cDNA are not able to bind bFGF. It is demonstrated that free heparin and heparan sulfate can reconstitute a low affinity receptor that is, in turn, required for the high affinity binding of bFGF. These studies suggest that the low affinity receptor is an accessory molecule required for binding of bFGF to the high affinity site. Such an obligatory interaction of low and high affinity FGF receptors suggests a physiological role for heparin-like, low affinity receptors and constitutes a novel mechanism for the regulation of growth factor-receptor interactions.  相似文献   

7.
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   

8.
Staphylococcus aureus, which mediated binding to heparan sulfate, and also strains of coagulase-negative staphylococci (CNS) adhered in high numbers to polymers with end-point attached heparin. A characteristic feature of several cell growth factors is strong affinity for heparin. In the present study, binding of the 125I-labeled heparin-binding growth factors (HBGF), acidic and basic fibroblast growth factor (aFGF, bFGF), and platelet-derived growth factor (PDGF) by S. aureus and CNS strains was examined. Staphylococcal strains used in this study bind bFGF and PDGF, but not aFGF. The binding of bFGF and PDGF was time dependent, influenced by pH and ionic strength for S. aureus Cowan 1. Preincubation of staphylococcal cells with unlabeled bFGF enhanced bFGF binding, but heparin, protamine sulfate, poly-L-lysine, and suramin were potent inhibitors of 125I-bFGF binding to cells of S. aureus Cowan 1. Glycosaminoglycans of comparable size (chondroitin sulfate), other polysulfated polymers (λ-carrageenan, fucoidan), and some polysulfated polysaccharides (dextran sulfate, pentosan polysulfate) inhibited binding of both GFs to various extents. The partial inhibition of binding of both GFs after protease and periodate treatments indicates that both proteinaceous and other carbohydrate moieties participate in the binding. A lysozyme cell surface extract and bacterial lysates of S. aureus Cowan 1 competitively inhibited binding of 125I-bFGF and 125I-PDGF. These results suggest that staphylococci have the ability to bind two of the HBGFs, bFGF and PDGF, but not aFGF, via more than one cell structure. These binding structures seem to be exposed on the cell surface and deeply anchored in the cytoplasmic membrane as well.  相似文献   

9.
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   

10.
Basic fibroblast growth factor (bFGF) is a known mitogen for vascular smooth muscle cells and has been implicated as having a role in a number of proliferative vascular disorders. Binding of bFGF to heparin or heparan sulfate has been demonstrated to both stimulate and inhibit growth factor activity. The activity, towards bFGF, of heparan sulfate proteoglycans present within the vascular system is likely related to the chemical characteristics of the glycosaminoglycan as well as the structure and pericellular location of the intact proteoglycans. We have previously shown that endothelial conditioned medium inhibits both bFGF binding to vascular smooth muscle cells and bFGF stimulated cell proliferation in vitro. In the present study, we have isolated proteoglycans from endothelial cell conditioned medium and demonstrated that they are responsible for the bFGF inhibitory activity. We further separated endothelial secreted proteoglycans into two fractions, PG-A and PG-B. The larger sized fraction (PG-A) had greater inhibitory activity than did PG-B for both bFGF binding and bFGF stimulation of vascular smooth muscle cell proliferation. The increased relative activity of PG-A was attributed, in part, to larger heparan sulfate chains which were more potent inhibitors of bFGF binding than the smaller heparan sulfate chains on PG-B. Both proteoglycan fractions contained perlecan-like core proteins; however, PG-A contained an additional core protein (approximately 190 kDa) that was not observed in PG-B. Both proteoglycan fractions bound bFGF directly, and PG-A bound a significantly greater relative amount of bFGF than did PG-B. Thus the ability of endothelial heparan sulfate proteoglycans to bind bFGF and prevent its association with vascular smooth muscle cells appears essential for inhibition of bFGF-induced mitogenesis. The production of potent bFGF inhibitory heparan sulfate proteoglycans by endothelial cells might contribute to the maintenance of vascular homeostasis. J. Cell. Physiol. 172:209–220, 1997. © 1997 Wiley-Liss, Inc.  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 PHEX gene (phosphate-regulating gene with homologies to endopeptidase on the X chromosome) identified as a mutated gene in patients with X-linked hypophosphatemia (XLH), encodes a protein (PHEX) that shows striking homologies to members of the M13 family of zinc metallopeptidases. In the present work the interaction of glycosaminoglycans with PHEX has been investigated by affinity chromatography, circular dichroism, protein intrinsic fluorescence analysis, hydrolysis of FRET substrates flow cytometry and confocal microscopy. PHEX was eluted from a heparin-Sepharose chromatography column at 0.8 M NaCl showing a strong interaction with heparin. Circular dichroism spectra and intrinsic fluorescence analysis showed that PHEX is protected by glycosaminoglycans against thermal denaturation. Heparin, heparan sulfate and chondroitin sulfate inhibited PHEX catalytic activity, however among them, heparin presented the highest inhibitory activity (Ki = 2.5 ± 0.2 nM). Flow cytometry analysis showed that PHEX conjugated to Alexa Fluor 488 binds to the cell surface of CHO-K1, but did not bind to glycosaminoglycans defective cells CHO-745. Endogenous PHEX was detected at the cell surface of CHO-K1 colocalized with heparan sulfate proteoglycans, but was not found at the cell surface of glycosaminoglycans defective cells CHO-745. In permeabilized cells, PHEX was detected in endoplasmic reticulum of both cells. In addition, we observed that PHEX colocalizes with heparan sulfate at the cell surface of osteoblasts. This is the first report that the metallopeptidase PHEX is a heparin binding protein and that the interaction with GAGs modulates its enzymatic activity, protein stability and cellular trafficking.  相似文献   

13.
Basic fibroblast growth factor (bFGF) is a therapeutic target of anti-angiogenesis. Here, we report that a novel sulfated glycopeptide derived from Gekko swinhonis Guenther (GSPP), an anticancer drug in traditional Chinese medicine, inhibits tumor angiogenesis by targeting bFGF. GSPP significantly decreased the production of bFGF in hepatoma cells by suppressing early growth response-1. GSPP inhibited the release of bFGF from extracellular matrix by blocking heparanase enzymatic activity. Moreover, GSPP competitively inhibited bFGF binding to heparin/heparan sulfate via direct binding to bFGF. Importantly, GSPP abrogated the bFGF-stimulated proliferation and migration of endothelial cells, whereas it had no inhibitory effect on endothelial cells in the absence of bFGF. Further study revealed that GSPP prevented bFGF-induced neovascularization and inhibited tumor angiogenesis and tumor growth in a xenograft mouse model. These results demonstrate that GSPP inhibits tumor angiogenesis by blocking bFGF production, release from the extracellular matrix, and binding to its low affinity receptor, heparin/heparan sulfate.  相似文献   

14.
We examined the ability of Chinese hamster ovary (CHO) cell mutants defective in glycosaminoglycan synthesis to metabolize 125I-labeled thrombospondin (TSP). Wild type CHO cells bound and degraded 125I-TSP with kinetics similar to those reported for endothelial cells. Both binding and degradation were saturable (half-saturation at 20 micrograms/ml). When the concentration of labeled TSP was 1-5 micrograms/ml, mutant 745, defective in xylosyltransferase, and mutant 761, defective in galactosyltransferase I, bound and degraded 6- to 16-fold less TSP than wild type; mutant 803, which specifically lacks heparan sulfate chains, bound and degraded 5-fold less TSP than wild type; and mutant 677, which lacks heparan sulfate and has increased levels of chondroitin sulfate, bound and degraded 2-fold less TSP than wild type. Binding and degradation of TSP by the mutants were not saturable at TSP concentrations up to 100 micrograms/ml. Bound TSP was localized by immunofluorescence to punctate structures on wild type and, to a lesser extent, 677 cells. Heparitinase pretreatment of wild type cells caused a 2- to 3-fold decrease in binding and degradation, whereas chondroitinase pretreatment had no effect. Chondroitinase pretreatment of the 677 mutant (deficient heparan sulfate and excess chondroitin sulfate) caused a 2-fold decrease in binding and an 8-fold decrease in turnover, whereas heparitinase pretreatment had no effect. Treatment of wild type cells with both heparitinase and chondroitinase resulted in a 6- to 8-fold decrease in binding and turnover. These results indicate that cell surface proteoglycans mediate metabolism of TSP by CHO cells and that the primary effectors of TSP metabolism are heparan sulfate proteoglycans.  相似文献   

15.
Promotion of cell growth and differentiation by growth factors during early development and organ formation are both temporally and spatially very precise. Syndecan is a well characterized integral membrane proteoglycan that binds several extracellular matrix components via its heparan sulfate chains and is therefore suggested to participate in cell regulation. Syndecan-like molecules, as low affinity receptors for heparin-binding growth factors, have been recently suggested to also regulate growth factor activity. Heparin/heparan sulfate interaction is required before, e.g. basic fibroblast growth factor (bFGF) can associate with its high affinity cell surface receptors and trigger signal transduction. In this paper we show that syndecan, but not free heparan sulfate chains, can simultaneously bind both bFGF and extracellular matrix molecules. Moreover, increased DNA synthesis of 3T3 cells was observed when the 3T3 cells were exposed to beads coated with the fibronectin-syndecan-bFGF complex, indicating that bFGF remains biologically active even when immobilized to matrix via the heparan sulfate chains of syndecan. Finally, when bFGF was bound to the surface of another cell type (epithelial), co-culture with 3T3 cells stimulated 3T3 cell growth. Therefore, we suggest that syndecan-like molecules may determine sites of growth factor action at cell-matrix and cell-cell interfaces.  相似文献   

16.
Deposition of β cell toxic islet amyloid is a cardinal finding in type 2 diabetes. In addition to the main amyloid component islet amyloid polypeptide (IAPP), heparan sulfate proteoglycan is constantly present in the amyloid deposit. Heparan sulfate (HS) side chains bind to IAPP, inducing conformational changes of the IAPP structure and an acceleration of fibril formation. We generated a double-transgenic mouse strain (hpa-hIAPP) that overexpresses human heparanase and human IAPP but is deficient of endogenous mouse IAPP. Culture of hpa-hIAPP islets in 20 mm glucose resulted in less amyloid formation compared with the amyloid load developed in cultured islets isolated from littermates expressing human IAPP only. A similar reduction of amyloid was achieved when human islets were cultured in the presence of heparin fragments. Furthermore, we used CHO cells and the mutant CHO pgsD-677 cell line (deficient in HS synthesis) to explore the effect of cellular HS on IAPP-induced cytotoxicity. Seeding of IAPP aggregation on CHO cells resulted in caspase-3 activation and apoptosis that could be prevented by inhibition of caspase-8. No IAPP-induced apoptosis was seen in HS-deficient CHO pgsD-677 cells. These results suggest that β cell death caused by extracellular IAPP requires membrane-bound HS. The interaction between HS and IAPP or the subsequent effects represent a possible therapeutic target whose blockage can lead to a prolonged survival of β cells.  相似文献   

17.
The seinile plaques found within the cerebral cortex and hippocampus of the Alzheimer disease brain contain β-amyloid peptide (Aβ) fibrils that are associated with a variety of macromolecular species, including dermatan sulfate proteoglycan and heparan sulfate proteoglycan. The latter has been shown recently to bind tightly to both amyloid precursor protein and A/β, and this binding has been attributed largely to the interaction of the core protein of heparan sulfate proteoglycan with Aβ and its precursor. Here we have examined the ability of synthetic Aβ s to bind to and interact with the glycosaminoglycan moieties of proteoglycans. Aβ(1–28) associates with heparin, heparan sulfate, dermatan sulfate, and chondroitin sulfate. The interaction of these sulfated polysaccharides with the amyloid peptide results in the formation of large aggregates that are readily sedimented by centrifugation. The ability of both Aβ(1–28) and Aβ(1–40) to bind glycosaminoglycans is pH-dependent, with increasing interaction as the pH values fall below neutrality and very little binding at pH 8.0. The pH profile of heparin-induced aggregation of Aβ(1–28) has a midpoint pH of approximately 6.5, suggesting that one or more histidine residues must be protonated for binding to occur. Analysis of the Aβ sequence reveals a consensus heparin-binding domain at residues 12–17, and this motif contains histidines at positions 13 and 14 that may be involved in the interaction with glycosaminoglycans. This hypothesis is supported by the following observations: (a) Aβ(13–17) binds tightly to a heparin affinity column at pH 4.0, but not at pH 8.0; and (b) an Aβ(13–17) in which histidine residues 13 and 14 have been replaced with serines does not bind to a heparin column at either pH 8.0 or 4.0. Together, the data indicate that Aβ is capable of binding to the glycosaminoglycan chains of proteoglycans, and such an interaction may be relevant to the etiology and pathology of Alzheimer's disease.  相似文献   

18.
Animal cells utilize beta-D-xylosides as primers for glycosaminoglycan synthesis. However, most xylosides preferentially stimulate chondroitin sulfate synthesis and only weakly prime heparan sulfate synthesis. To test if the structure of the aglycone determines the type of glycosaminoglycan made, the priming activity of methyl, n-octyl, p-nitrophenyl, 4-methylumbelliferyl, trans,trans-farnesyl, cholesteryl, and estradiol beta-D-xylosides was compared. Their potency was tested in pgsA-745 cells, a Chinese hamster ovary cell mutant unable to initiate glycosaminoglycan synthesis due to a defect in xylosyltransferase. All of the xylosides stimulated chondroitin sulfate synthesis in the mutant, but only estradiol beta-D-xyloside primed heparan sulfate synthesis efficiently. When incubated with 30 microM estradiol beta-D-xyloside, mutant cells made about 3-fold more glycosaminoglycan than untreated wild-type cells and as much as 50% was heparan sulfate. Estradiol beta-D-xyloside also induced heparan sulfate synthesis in cycloheximide-treated wild-type Chinese hamster ovary cells, bovine aortic endothelial cells, baby hamster kidney cells, and Balb/c 3T3 fibroblasts. In addition to stimulating heparan sulfate synthesis, low concentrations of estradiol beta-D-xyloside inhibited the formation of endogenous heparan sulfate proteoglycans.  相似文献   

19.
Cathepsin X binds to cell surface heparan sulfate proteoglycans   总被引:3,自引:0,他引:3  
Glycosaminoglycans have been shown to be important regulators of activity of several papain-like cathepsins. Binding of glycosaminoglycans to cathepsins thus directly affects catalytic activity, stability or the rate of autocatalytic activation of cathepsins. The interaction between cathepsin X and heparin has been revealed by affinity chromatography using heparin-Sepharose. Conformational changes were observed to accompany heparin-cathepsin X interaction by far UV-circular dichroism at both acidic (4.5) and neutral (7.4) pH. These conformational changes promoted a 4-fold increase in the dissociation constant of the enzyme-substrate interaction and increased 2.6-fold the kcat value also. The interaction between cathepsin X and heparin or heparan sulfate is specific since dermatan sulfate, chondroitin sulfate, and hyaluronic acid had no effect on the cathepsin X activity. Using flow cytometry cathepsin X was shown to bind cell surface heparan sulfate proteoglycans in wild-type CHO cells but not in CHO-745 cells, which are deficient in glycosaminoglycan synthesis. Moreover, fluorescently labeled cathepsin X was shown by confocal microscopy to be endocytosed by wild-type CHO cells, but not by CHO-745 cells. These results demonstrate the existence of an endocytosis mechanism of cathepsin X by the CHO cells dependent on heparan sulfate proteoglycans present at the cell surface, thus strongly suggesting that heparan sulfate proteoglycans can regulate the cellular trafficking and the enzymatic activity of cathepsin X.  相似文献   

20.
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   

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