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 46 毫秒
1.
Copper-dependent co-internalization of the prion protein and glypican-1   总被引:1,自引:0,他引:1  
Heparan sulfate chains have been found to be associated with amyloid deposits in a number of diseases including transmissible spongiform encephalopathies. Diverse lines of evidence have linked proteoglycans and their glycosaminoglycan chains, and especially heparan sulfate, to the metabolism of the prion protein isoforms. Glypicans are a family of glycosylphosphatidylinositol-anchored, heparan sulfate-containing, cell-associated proteoglycans. Cysteines in glypican-1 can become nitrosylated by endogenously produced nitric oxide. When glypican-1 is exposed to a reducing agent, such as ascorbate, nitric oxide is released and autocatalyses deaminative cleavage of heparan sulfate chains. These processes take place while glypican-1 recycles via a non-classical, caveolin-associated pathway. We have previously demonstrated that prion protein provides the Cu2+ ions required to nitrosylate thiol groups in the core protein of glypican-1. By using confocal immunofluorescence microscopy and immunomagnetic techniques, we now show that copper induces co-internalization of prion protein and glypican-1 from the cell surface to perinuclear compartments. We find that prion protein is controlling both the internalization of glypican-1 and its nitric oxide-dependent autoprocessing. Silencing glypican-1 expression has no effect on copper-stimulated prion protein endocytosis, but in cells expressing a prion protein construct lacking the copper binding domain internalization of glypican-1 is much reduced and autoprocessing is abrogated. We also demonstrate that heparan sulfate chains of glypican-1 are poorly degraded in prion null fibroblasts. The addition of either Cu2+ ions, nitric oxide donors, ascorbate or ectopic expression of prion protein restores heparan sulfate degradation. These results indicate that the interaction between glypican-1 and Cu2+-loaded prion protein is required both for co-internalization and glypican-1 self-pruning.  相似文献   

2.
VEGF165 binding to endothelial cells is potentiated by glycosaminoglycans (GAGs). Here, we have investigated the impact of VEGF165 N-glycosylation on GAG binding. Although glycosylated VEGF165 bound to heparin with only slightly higher affinity than non-glycosylated VEGF165, the natural ligand heparan sulfate induced a conformational change only in the glycosylated protein. Unfolding studies of the VEGF proteins indicated a stabilising effect of heparin on the growth factor structure.  相似文献   

3.
Fibrillins form multifunctional microfibrils in most connective tissues. Deficiencies in fibrillin assembly can result in fibrillinopathies, such as Marfan syndrome. We demonstrate the presence of heparin/heparan sulfate binding sites in fibrillin-2 and -3. Multimerization of all three fibrillins drastically increased the apparent affinity of their interaction with heparin/heparan sulfate. Surprisingly, contrary to other reports heparin/heparan sulfate strongly inhibited homo- and heterotypic N-to-C-terminal fibrillin interactions. These data suggest that heparin/heparan sulfate controls the formation of microfibrils at the bead interaction stage.  相似文献   

4.
Herndon  ME; Stipp  CS; Lander  AD 《Glycobiology》1999,9(2):143-155
The method of affinity coelectrophoresis was used to study the binding of nine representative glycosaminoglycan (GAG)-binding proteins, all thought to play roles in nervous system development, to GAGs and proteoglycans isolated from developing rat brain. Binding to heparin and non-neural heparan and chondroitin sulfates was also measured. All nine proteins-laminin-1, fibronectin, thrombospondin-1, NCAM, L1, protease nexin-1, urokinase plasminogen activator, thrombin, and fibroblast growth factor-2-bound brain heparan sulfate less strongly than heparin, but the degree of difference in affinity varied considerably. Protease nexin-1 bound brain heparan sulfate only 1.8- fold less tightly than heparin (Kdvalues of 35 vs. 20 nM, respectively), whereas NCAM and L1 bound heparin well (Kd approximately 140 nM) but failed to bind detectably to brain heparan sulfate (Kd>3 microM). Four proteins bound brain chondroitin sulfate, with affinities equal to or a few fold stronger than the same proteins displayed toward cartilage chondroitin sulfate. Overall, the highest affinities were observed with intact heparan sulfate proteoglycans: laminin-1's affinities for the proteoglycans cerebroglycan (glypican-2), glypican-1 and syndecan-3 were 300- to 1800-fold stronger than its affinity for brain heparan sulfate. In contrast, the affinities of fibroblast growth factor-2 for cerebroglycan and for brain heparan sulfate were similar. Interestingly, partial proteolysis of cerebroglycan resulted in a >400- fold loss of laminin affinity. These data support the views that (1) GAG-binding proteins can be differentially sensitive to variations in GAG structure, and (2) core proteins can have dramatic, ligand-specific influences on protein-proteoglycan interactions.   相似文献   

5.
Abstract The pathogenic bacterium Helicobacter pylori , which causes active, chronic type B gastritis and peptic ulcer disease, and increases the risk for development of gastric cancer, could tentatively interfere with growth factors and growth factor receptors of importance for the gastroduodenal mucosa, e.g. heparin-binding FGFs (fibroblast growth factors). H. pylori binds FGF with an extremely strong affinity (3.8 × 10−12 M), and also heparan sulfate and heparin with higher affinity ( K d 9 × 10−9 M) than FGFs bind to heparin (10−8–10−9 M). FGF receptors are also dependent on heparin for their activation. Heparan sulfate binding proteins (HSBP) are exposed on and shed from the surface of H. pylori , which often are localised close to the epithelial stem cells in the gastroduodenal glands. H. pylori could thus efficiently interfere with growth factors and growth factor receptors, tentatively resulting in disturbance of the delicate balance that control the renewal, maintenance and repair of the gastroduodenal mucosa. This mode of action has previously not been considered, but may constitute part of its pathogenic mechanism. Such a dynamic mode of action of H. pylori may explain the reason for that infected victims may either suffer from gastrointestinal symptoms or lack clinical evidence of disease or discomfort.  相似文献   

6.
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   

7.
We have previously demonstrated that the Slit proteins, which are involved in axonal guidance and related processes, are high-affinity ligands of the heparan sulfate proteoglycan glypican-1. Glypican-Slit protein interactions have now been characterized in greater detail using two approaches. The ability of heparin oligosaccharides of defined structure (ranging in size from disaccharide to tetradeccasaccharide) to inhibit binding of a glypican-Fc fusion protein to recombinant human Slit-2 was determined using an ELISA. Surface plasmon resonance (SPR) spectroscopy, which measures the interactions in real time, was applied for quantitative modeling of heparin-Slit binding on heparin biochips. Heparin was covalently immobilized on these chips through a pre-formed albumin-heparin conjugate, and the inhibition of Slit binding by heparin, LMW heparin, and heparin-derived oligosaccharides (di-, tetra-, hexa-, and octa-) was examined utilizing solution competition SPR. These competition studies demonstrate that the smallest heparin oligosaccharide competing with heparin binding to Slit was a tetrasaccharide, and that in the ELISA maximum inhibition (approximately 60% at 2 microM concentration) was attained with a dodecasaccharide.  相似文献   

8.
The monoclonal antibody 10E4, which recognizes an epitope supposed to contain N-unsubstituted glucosamine, is commonly used to trace heparan sulfate proteoglycans. It has not been fully clarified if the N-unsubstituted glucosamine is required for antibody recognition and if all heparan sulfates carry this epitope. Here we show that the epitope can contain N-unsubstituted glucosamine and that nitric oxide-generated deaminative cleavage at this residue in vivo can destroy the epitope. Studies using flow cytometry and confocal immunofluorescence microscopy of both normal and transformed cells indicated that the 10E4 epitope was partially inaccessible in the heparan sulfate chains attached to glypican-1. The 10E4 antibody recognized mainly heparan sulfate degradation products that colocalized with acidic endosomes. These sites were greatly depleted of 10E4-positive heparan sulfate on suramin inhibition of heparanase. Instead, there was increased colocalization between 10E4-positive heparan sulfate and glypican-1. When both S-nitrosylation of Gpc-1 and heparanase were inhibited, detectable 10E4 epitope colocalized entirely with glypican-1. In nitric oxide-depleted cells, there was both an increased signal from 10E4 and increased colocalization with glypican-1. In suramin-treated cells, the 10E4 epitope was destroyed by ascorbate-released nitric oxide with concomitant formation of anhydromannose-containing heparan sulfate oligosaccharides. Immunoisolation of radiolabeled 10E4-positive material from unperturbed cells yielded very little glypican-1 when compared with specifically immunoisolated glypican-1 and total proteoglycan and degradation products. The 10E4 immunoisolates were either other heparan sulfate proteoglycans or heparan sulfate degradation products.  相似文献   

9.
Structural Properties of the Heparan Sulfate Proteoglycans of Brain   总被引:1,自引:1,他引:0  
The heparan sulfate proteoglycans present in a deoxycholate extract of rat brain were purified by ion exchange chromatography, affinity chromatography on lipoprotein lipase agarose, and gel filtration. Heparitinase treatment of the heparan sulfate proteoglycan fraction (containing 86% heparan sulfate and 10% chondroitin sulfate) that was eluted from the lipoprotein lipase affinity column with 1 M NaCl led to the appearance of a major protein core with a molecular size of 55,000 daltons, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Comparison of the effects of heparinase and heparitinase treatment revealed that the heparan sulfate proteoglycans of brain contain a significant proportion of relatively short N-sulfoglucosaminyl 6-O-sulfate [or N-sulfoglucosaminyl](alpha 1-4)iduronosyl 2-O-sulfate(alpha 1-4) repeating units and that the portions of the heparan sulfate chains in the vicinity of the carbohydrate-protein linkage region are characterized by the presence of D-glucuronic acid rather than L-iduronic acid. After chondroitinase treatment of a proteoglycan fraction that contained 62% chondroitin sulfate and 21% heparan sulfate (eluted from lipoprotein lipase with 0.4 M NaCl), the charge and density of a portion of the heparan sulfate-containing proteoglycans decreased significantly. These results indicate that a population of "hybrid" brain proteoglycans exists that contain both chondroitin sulfate and heparan sulfate chains covalently linked to a common protein core.  相似文献   

10.
The platelet-derived growth factor (PDGF) family comprises disulfide-bonded dimeric isoforms and plays a key role in the proliferation and migration of mesenchymal cells. Traditionally, it consists of homo- and heterodimers of A and B polypeptide chains that occur as long (AL and BL) or short (AS and BS) isoforms. Short isoforms lack the basic C-terminal extension that mediates binding to heparin. In the present study, we show that certain PDGF isoforms bind in a specific manner to glycosaminoglycans (GAGs). Experiments performed with wild-type and mutant Chinese hamster ovary cells deficient in the synthesis of GAGs revealed that PDGF long isoforms bind to heparan sulfate and chondroitin sulfate, while PDGF short isoforms only bind to heparan sulfate. This was confirmed by digestion of cell surface GAGs with heparitinase and chondroitinase ABC and by incubation with sodium chloride to prevent GAG sulfation. Furthermore, exogenous GAGs inhibited the binding of long isoforms to the cell membrane more efficiently than that of short isoforms. Additionally, we performed surface plasmon resonance experiments to study the inhibition of PDGF isoforms binding to low molecular weight heparin by GAGs. These experiments showed that PDGF-AAL and PDGF-BBS isoforms bound to GAGs with the highest affinity. In conclusion, PDGF activity at the cell surface may depend on the expression of various cellular GAG species.  相似文献   

11.
Heparin binding by the HIV-1 tat protein transduction domain   总被引:5,自引:0,他引:5       下载免费PDF全文
The protein transduction domain from the HIV-1 tat protein (termed PTD-tat) has been fused to the C-terminus of a model cargo protein, the IgG binding domain of streptococcal protein G. We demonstrate that PG-Ctat (PTD-tat fused to the C-terminus of protein G) binds to a heparin affinity column. PG-Ctat binds with relatively high affinity, as shown by its elution at 1.6 M NaCl. The heparin binding properties of PTD-tat are consistent with the idea that heparan sulfate, an analog of heparin found at the cell surface, plays a role in the translocation of PTD-tat fusions. We suggest that the heparin-binding properties of PTD-tat can be exploited for purification of PTD-tat fusions in the absence of affinity tags.  相似文献   

12.
Burkitt's lymphoma (BL) is a B-cell malignancy associated with the Epstein-Barr virus (EBV). Mounting evidence has implicated heparan sulfate proteoglycans and heparan sulfate-like glycosaminoglycans (HSGAGs) in the initiation, severity, and progression of the malignancy. The importance of HSGAGs in regulating BL cell growth was therefore examined. Extracellular exogenous heparin inhibited cell growth >30%, while heparin internalized with poly(beta-amino ester)s promoted proliferation up to 58%. The growth-modulating effects of heparin and internalized heparin were dependent on cell surface HSGAGs, PI3K, and Erk/Mek. Treatment of cells with protamine sulfate or with heparinases potently inhibited proliferation, with the greatest effects induced by heparinase I. Cell surface HSGAGs therefore play an important role in regulating BL proliferation and may offer a potential target for therapeutic intervention.  相似文献   

13.
The adipokine tartrate resistant acid phosphatase (TRAP) 5a isoform exerts a growth factor-effect on pre-adipocytes. This study aimed to identify potential TRAP 5a interacting proteins in pre-adipocytes using pull down assays in combination with mass spectrometry. Nidogen-2, a protein shown to be expressed intracellularly and for secretion by pre-adipocytes, was shown to interact, through its globular G3 domain, with TRAP 5a in vitro. In vivo, TRAP 5a interacted with nidogen-2 in cultured 3T3-L1 mouse pre-adipocytes, as well as with transforming growth factor-β (TGF-β) interacting protein (TRIP-1), which is a protein that has previously been suggested to interact with TRAP in bone. In addition, TRAP 5a and nidogen-2 co-localized in adipose tissue cells in situ. These results indicate that TRAP 5a interacts with nidogen-2 and TRIP-1 in pre-adipocytic cells.  相似文献   

14.
Binding of heparin/heparan sulfate to fibroblast growth factor receptor 4   总被引:4,自引:0,他引:4  
Fibroblast growth factors (FGFs) are heparin-binding polypeptides that affect the growth, differentiation, and migration of many cell types. FGFs signal by binding and activating cell surface FGF receptors (FGFRs) with intracellular tyrosine kinase domains. The signaling involves ligand-induced receptor dimerization and autophosphorylation, followed by downstream transfer of the signal. The sulfated glycosaminoglycans heparin and heparan sulfate bind both FGFs and FGFRs and enhance FGF signaling by mediating complex formation between the growth factor and receptor components. Whereas the heparin/heparan sulfate structures involved in FGF binding have been studied in some detail, little information has been available on saccharide structures mediating binding to FGFRs. We have performed structural characterization of heparin/heparan sulfate oligosaccharides with affinity toward FGFR4. The binding of heparin oligosaccharides to FGFR4 increased with increasing fragment length, the minimal binding domains being contained within eight monosaccharide units. The FGFR4-binding saccharide domains contained both 2-O-sulfated iduronic acid and 6-O-sulfated N-sulfoglucosamine residues, as shown by experiments with selectively desulfated heparin, compositional disaccharide analysis, and a novel exoenzyme-based sequence analysis of heparan sulfate oligosaccharides. Structurally distinct heparan sulfate octasaccharides differed in binding to FGFR4. Sequence analysis suggested that the affinity of the interaction depended on the number of 6-O-sulfate groups but not on their precise location.  相似文献   

15.
Endostatin (20 kDa) is a C-terminal proteolytic fragment of collagen XVIII that is localized in vascular basement membrane zones in various organs. It binds zinc, heparin/heparan sulfate, laminin, and sulfatides and inhibits angiogenesis and tumor growth. Here we determined the kinetics and affinity of the interaction of endostatin with heparin/heparan sulfate and investigated the effects of divalent cations on these interactions and on the biological activities of endostatin. The binding of human recombinant endostatin to heparin and heparan sulfate was studied by surface plasmon resonance using BIAcore technology and further characterized by docking and molecular dynamics simulations. Kinetic data, evaluated using a 1:1 interaction model, showed that heparan sulfate bound to and dissociated from endostatin faster than heparin and that endostatin bound to heparin and heparan sulfate with a moderate affinity (K(D) approximately 2 microm). Molecular modeling of the complex between endostatin and heparin oligosaccharides predicted that, compared with mutagenesis studies, two further arginine residues, Arg(47) and Arg(66), participated in the binding. The binding of endostatin to heparin and heparan sulfate required the presence of divalent cations. The addition of ZnCl(2) to endostatin enhanced its binding to heparan sulfate by approximately 40% as well as its antiproliferative effect on endothelial cells stimulated by fibroblast growth factor-2, suggesting that this activity is mediated by the binding of endostatin to heparan sulfate. In contrast, no increase in the antiangiogenic and anti-proliferative activities of endostatin promoted by vascular endothelial growth factor was observed upon the addition of zinc.  相似文献   

16.
Abstract: We have characterized the structural properties of heparan sulfates from brain and other tissues after de-polymerization with a mixture of three heparin and heparan sulfate lyases from Flavobacterium heparinum. The resulting disaccharides were separated by HPLC and identified by comparison with authentic standards. In rat, rabbit, and bovine brain, 46–69% of the heparan sulfate disaccharides are N-acetylated and unsulfated, and 17–21% contain a single sulfate residue in the form of a sulfoamino group. In rabbit, bovine, and 1-day postnatal rat brain, disaccharides containing both a sulfated uronic acid and N-sulfate account for an additional 10–14%, together with smaller and approximately equall proportions (5–9%) of mono-, di-, and trisulfated disaccharides having sulfate at the 6-position of the glucosamine residue. Kidney and lung heparan sulfates are distinguished by high concentrations of disaccharides containing 6-sulfated N-acetylglucosamine residues. In chromaffin granules, the catecholamine-and peptide-storing organelles of adrenal medulla, where heparan sulfate accounts for a minor portion (5–10%) of the glycosaminoglycans, we have determined that bovine chromaffin granule membranes contain heparan sulfate in which almost all of the disaccharides are either unsulfated (71 %) or monosulfated (18%). In sympathetic nerves, norepinephrine is stored in large densecored vesicles that in biochemical composition and properties closely resemble adrenal chromaffin granules. However, in contrast to chromaffin granules, heparan sulfate accounts for ~ 75% of the total glycosaminoglycans in large dense-cored vesicles and more closely resembles heparin, insofar as it contains only 21 % unsulfated disaccharides, 10% mono-and disulfated disaccharides, and 69% trisulfated disaccharides. Our results therefore reveal significant differences among heparan sulfates from different sources, supporting other evidence that structural variations in heparan sulfate may be related to specific biological functions, such as the switching in the neural response from fibroblast growth factor-2 to fibro-blast growth factor-1 resulting from developmental changes in the glycosaminoglycan chains of a heparan sulfate proteoglycan.  相似文献   

17.
Heparan sulfate proteoglycans are critical binding partners for extracellular tranglutaminase-2 (TG2), a multifunctional protein involved in tissue remodeling events related to organ fibrosis and cancer progression. We previously showed that TG2 has a strong affinity for heparan sulfate (HS)/heparin and reported that the heparan sulfate proteoglycan syndecan-4 acts as a receptor for TG2 via its HS chains in two ways: by increasing TG2-cell surface trafficking/externalization and by mediating RGD-independent cell adhesion to fibronectin-TG2 matrix during wound healing. Here we have investigated the molecular basis of this interaction. Site-directed mutagenesis revealed that either mutation of basic RRWK (262-265) or KQKRK (598-602) clusters, forming accessible heparin binding sequences on the TG2 three-dimensional structure, led to an almost complete reduction of heparin binding, indicating that both clusters contribute to form a single binding surface. Mutation of residues Arg(19) and Arg(28) also led to a significant reduction in heparin binding, suggesting their involvement. Our findings indicate that the heparin binding sites on TG2 mainly comprise two clusters of basic amino acids, which are distant in the linear sequence but brought into spatial proximity in the folded "closed" protein, forming a high affinity heparin binding site. Molecular modeling showed that the identified site can make contact with a single heparin-derived pentasaccharide. The TG2-heparin binding mutants supported only weak RGD-independent cell adhesion compared with wild type TG2 or mutants with retained heparin binding, and both heparin binding clusters were critical for TG2-mediated cell adhesion. These findings significantly advance our knowledge of how HS/heparin influences the adhesive function of TG2.  相似文献   

18.
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   

19.
Growth and remodeling of lymphatic vasculature occur during development and during various pathologic states. A major stimulus for this process is the unique lymphatic vascular endothelial growth factor-C (VEGF-C). Other endothelial growth factors, such as fibroblast growth factor-2 (FGF-2) or VEGF-A, may also contribute. Heparan sulfate is a linear sulfated polysaccharide that facilitates binding and action of some vascular growth factors such as FGF-2 and VEGF-A. However, a direct role for heparan sulfate in lymphatic endothelial growth and sprouting responses, including those mediated by VEGF-C, remains to be examined. We demonstrate that VEGF-C binds to heparan sulfate purified from primary lymphatic endothelia, and activation of lymphatic endothelial Erk1/2 in response to VEGF-C is reduced by interference with heparin or pretreatment of cells with heparinase, which destroys heparan sulfate. Such treatment also inhibited phosphorylation of the major VEGF-C receptor VEGFR-3 upon VEGF-C stimulation. Silencing lymphatic heparan sulfate chain biosynthesis inhibited VEGF-C-mediated Erk1/2 activation and abrogated VEGFR-3 receptor-dependent binding of VEGF-C to the lymphatic endothelial surface. These findings prompted targeting of lymphatic N-deacetylase/N-sulfotransferase-1 (Ndst1), a major sulfate-modifying heparan sulfate biosynthetic enzyme. VEGF-C-mediated Erk1/2 phosphorylation was inhibited in Ndst1-silenced lymphatic endothelia, and scratch-assay responses to VEGF-C and FGF-2 were reduced in Ndst1-deficient cells. In addition, lymphatic Ndst1 deficiency abrogated cell-based growth and proliferation responses to VEGF-C. In other studies, lymphatic endothelia cultured ex vivo from Ndst1 gene-targeted mice demonstrated reduced VEGF-C- and FGF-2-mediated sprouting in collagen matrix. Lymphatic heparan sulfate may represent a novel molecular target for therapeutic intervention.  相似文献   

20.
Glypican-1 is a member of a family of glycosylphosphatidylinositol anchored cell surface heparan sulfate proteoglycans implicated in the control of cellular growth and differentiation. The 165-amino acid form of vascular endothelial growth factor (VEGF165) is a mitogen for endothelial cells and a potent angiogenic factor in vivo. Heparin binds to VEGF165 and enhances its binding to VEGF receptors. However, native HSPGs that bind VEGF165 and modulate its receptor binding have not been identified. Among the glypicans, glypican-1 is the only member that is expressed in the vascular system. We have therefore examined whether glypican-1 can interact with VEGF165. Glypican-1 from rat myoblasts binds specifically to VEGF165 but not to VEGF121. The binding has an apparent dissociation constant of 3 x 10(-10) M. The binding of glypican-1 to VEGF165 is mediated by the heparan sulfate chains of glypican-1, because heparinase treatment abolishes this interaction. Only an excess of heparin or heparan sulfates but not other types of glycosaminoglycans inhibited this interaction. VEGF165 interacts specifically not only with rat myoblast glypican-1 but also with human endothelial cell-derived glypican-1. The binding of 125I-VEGF165 to heparinase-treated human vascular endothelial cells is reduced following heparinase treatment, and addition of glypican-1 restores the binding. Glypican-1 also potentiates the binding of 125I-VEGF165 to a soluble extracellular domain of the VEGF receptor KDR/flk-1. Furthermore, we show that glypican-1 acts as an extracellular chaperone that can restore the receptor binding ability of VEGF165, which has been damaged by oxidation. Taken together, these results suggest that glypican-1 may play an important role in the control of angiogenesis by regulating the activity of VEGF165, a regulation that may be critical under conditions such as wound repair, in which oxidizing agents that can impair the activity of VEGF are produced, and in situations were the concentrations of active VEGF are limiting.  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