首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Heparin and heparan sulfate binding sites on B-16 melanoma cells   总被引:2,自引:0,他引:2  
We have reported previously that the production of a tumor cell factor that stimulates synthesis of fibroblast collagenase is influenced by a fibroblast-deposited matrix component, possibly heparan sulfate-proteoglycan. In this study, binding sites for heparin and heparan sulfate on mouse B-16 melanoma cells have been demonstrated. Binding of 3H-heparin and 35S-heparan sulfate has been shown to occur to whole cells, isolated membranes, and to a component(s) of detergent extracts of the membranes. Scatchard analysis of binding of 3H-heparin yielded a Kd of 2-5 x 10(-8) M and a Bmax of 0.5 x 10(7) heparin molecules bound per cell. Binding of 35S-heparan sulfate was of at least an order of magnitude lower affinity than heparin, but the Bmax was similar to that for heparin. Competition studies showed that 35S-heparan sulfate binding was inhibited totally by heparin and heparan sulfate and partially by dermatan sulfate, but no inhibition was obtained with hyaluronate or chondroitin sulfate. Binding of 3H-heparin was inhibited totally by heparin but to different extents by preparations of heparan sulfate from different tissue sources. The heparin/heparan sulfate binding activity is a protein(s) because it is destroyed by treatment with trypsin. Binding of 3H-heparin to transblots of the detergent extract of the B-16 cell membranes indicated that at least part of the binding activity is a 14,000-dalton protein.  相似文献   

2.
The binding of iodinated basic fibroblast growth factor (bFGF) to low-density heparan sulfate proteoglycan purified from the Engelbreth Holm Swarm (EHS) sarcoma was investigated using different techniques. The tumor clearly contained bFGF, the level being comparable to that found in other tissues such as human or bovine brain. 125I bFGF strongly bound to the basement membrane-like matrix of EHS frozen sections as revealed by autoradiography. Iodinated bFGF bound to purified heparan sulfate proteoglycan but not to laminin or collagen type IV, three components isolated from the same tumor. In contrast, acidic fibroblast growth factor (aFGF) displayed negligible binding to heparan sulfate proteoglycan. Binding of bFGF to frozen sections and to purified proteoglycan could be strongly inhibited by heparin and was displaced by an excess of unlabeled factor and completely suppressed after heparitinase and heparinase treatments. Binding was a function of the salt concentration and was abolished at 0.6 M NaCl. Scatchard analysis indicated the affinity site had a Kd of about 30 nM, a value 10-15 higher than that recently reported by Moscatelli (J. Cell. Physiol., 131:123-130, 1987) in the case of the low-affinity binding sites present on the surface of baby hamster kidney (BHK) cells.  相似文献   

3.
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   

4.
Foot-and-mouth disease virus (FMDV) enters cells by attaching to cellular receptor molecules of the integrin family, one of which has been identified as the RGD-binding integrin alpha(v)beta3. Here we report that, in addition to an integrin binding site, type O strains of FMDV share with natural ligands of alpha(v)beta3 (i.e., vitronectin and fibronectin) a specific affinity for heparin and that binding to the cellular form of this sulfated glycan, heparan sulfate, is required for efficient infection of cells in culture. Binding of the virus to paraformaldehyde-fixed cells was powerfully inhibited by agents such as heparin, that compete with heparan sulfate or by agents that compete for heparan sulfate (platelet factor 4) or that inactivate it (heparinase). Neither chondroitin sulfate, a structurally related component of the extracellular matrix, nor dextran sulfate appreciably inhibited binding. The functional importance of heparan sulfate binding was demonstrated by the facts that (i) infection of live cells by FMDV could also be blocked specifically by heparin, albeit at a much higher concentration of inhibitor; (ii) pretreatment of cells with heparinase reduced the number of plaques formed compared with that for untreated cells; and (iii) mutant cell lines deficient in heparan sulfate expression were unable to support plaque formation by FMDV, even though they remained equally susceptible to another picornavirus, bovine enterovirus. The results show that entry of type O FMDV into cells is a complex process and suggest that the initial contact with the cell surface is made through heparan sulfate.  相似文献   

5.
Scatchard analysis of binding of 125I-basic fibroblast growth factor (FGF) to baby hamster kidney (BHK) cells revealed the presence of two binding sites: a high affinity site with KD of 20 pM and 80,000 sites per cell and a low affinity site with KD of about 2 nM and 600,000 sites per cell. The binding to the two sites could be separated by first washing the cells with 2 M NaCl at pH 7.5 which released the low affinity binding and then extracting the cells with 0.5% Triton X-100 to recover the 125I-basic FGF bound to high affinity sites. The binding to the high affinity site was acid sensitive, suggesting that it represented binding to the receptor. Binding to the low affinity site could be competed strongly by heparin and less strongly by heparan sulfate but not by chondroitin sulfate, dermatan sulfate, or keratan sulfate. Treatment of BHK cells with heparinase abolished 62% of the low affinity binding, suggesting that the low affinity binding represented binding to cell-associated, heparin-like molecules. A variety of other cell types, including bovine capillary endothelial (BCE) cells, also demonstrated both low and high affinity binding sites. To test whether the low affinity binding might play a role in the basic FGF stimulation of plasminogen activator (PA) production by BCE cells, heparin was added to BCE cultures at concentrations which totally blocked binding of 125I-basic FGF to the low affinity sites. Addition of the heparin did not diminish the increased PA production induced by basic FGF. This suggests that the low affinity binding has no direct role in the stimulation of PA production in BCE cells.  相似文献   

6.
Follistatin, an activin-binding protein secreted by cultured rat granulosa cells, was shown to associate with the cell surface by affinity labeling with 125I-activin. Addition of follistatin to the cultured cells demonstrated a typical ligand-binding saturation curve, suggesting that granulosa cells have a specific binding site for follistatin. This binding was markedly inhibited by heparin and heparan sulfate, but not by chondroitin sulfates A and C, keratan sulfate, and dermatan sulfate. When granulosa cells were treated with glycosaminoglycan-degrading enzymes before or after addition of follistatin to the cultures, heparinase and heparitinase treatments resulted in significant suppression of the binding, whereas treatment with chondroitinase ABC had no effect. A competition study of the binding using heparin derivatives demonstrated that follistatin seemed to recognize O-sulfate groups in the heparin molecule. Heparitinase-treated granulosa cells exhibited almost full responsiveness to activin, indicating that the enzyme treatment had no effect on activin and receptor interaction. These results suggest that follistatin/activin-binding protein binds to heparan sulfate side chains of proteoglycans on the granulosa cell surface to regulate the various actions of activin.  相似文献   

7.
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   

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

9.
Heparin affin regulatory peptide (HARP) is a polypeptide belonging to a family of heparin binding growth/differentiation factors. The high affinity of HARP for heparin suggests that this secreted polypeptide should also bind to heparan sulfate proteoglycans derived from cell surface and extracellular matrix defined as extracellular compartments. Using Western blot analysis, we detected HARP bound to heparan sulfate proteoglycans in the extracellular compartments of MDA-MB 231 and MC 3T3-E1 as well as NIH3T3 cells overexpressing HARP protein. Heparitinase treatment of BEL cells inhibited HARP-induced cell proliferation, and the biological activity of HARP in this system was restored by the addition of heparin. We report that heparan sulfate, dermatan sulfate, and to a lesser extent, chondroitin sulfate A, displaced HARP bound to the extracellular compartment. Binding analyses with a biosensor showed that HARP bound heparin with fast association and dissociation kinetics (kass = 1.6 x 10(6) M-1 s-1; kdiss = 0.02 s-1), yielding a Kd value of 13 nM; the interaction between HARP and dermatan sulfate was characterized by slower association kinetics (kass = 0.68 x 10(6) M-1 s-1) and a lower affinity (Kd = 51 nM). Exogenous heparin, heparan sulfate, and dermatan sulfate potentiated the growth-stimulatory activity of HARP, suggesting that corresponding proteoglycans could be involved in the regulation of the mitogenic activity of HARP.  相似文献   

10.
TIMP-3 binds to sulfated glycosaminoglycans of the extracellular matrix   总被引:17,自引:0,他引:17  
Of the four known tissue inhibitors of metalloproteinases (TIMPs), TIMP-3 is distinguished by its tighter binding to the extracellular matrix. The present results show that glycosaminoglycans such as heparin, heparan sulfate, chondroitin sulfates A, B, and C, and sulfated compounds such as suramin and pentosan efficiently extract TIMP-3 from the postpartum rat uterus. Enzymatic treatment by heparinase III or chondroitinase ABC also releases TIMP-3, but neither one alone gives complete release. Confocal microscopy shows colocalization of heparan sulfate and TIMP-3 in the endometrium subjacent to the lumen of the uterus. Immunostaining of TIMP-3 is lost upon digestion of tissue sections with heparinase III and chondroitinase ABC. The N-terminal domain of human TIMP-3 was expressed and found to bind to heparin with affinity similar to that of full-length mouse TIMP-3. The A and B beta-strands of the N-terminal domain of TIMP-3 contain two potential heparin-binding sequences rich in lysine and arginine; these strands should form a double track on the outer surface of TIMP-3. Synthetic peptides corresponding to segments of these two strands compete for heparin in the DNase II binding assay. TIMP-3 binding may be important for the cellular regulation of activity of the matrix metalloproteinases.  相似文献   

11.
Binding and endocytosis of heparin by human endothelial cells in culture   总被引:8,自引:0,他引:8  
Binding of heparin and low molecular weight heparin fragments (CY 222, Mr range 1500-8000) to human vascular endothelial cells was studied. Primary culture of human umbilical vein endothelial cells and either 125I or 3H-labeled heparin or [125I]CY 222 were used. Slow, saturable and specific binding was found. No other tested glycosaminoglycan, excepting a highly sulfated heparan fraction, was able to compete for heparin binding. Two groups of binding sites for [3H]heparin could be distinguished: one with high affinity (Kd = 0.12 microM) and another with lower affinity (Kd = 1.37 microM) and a relative large capacity of binding (1.16 X 10(7) molecules/cell) was calculated. The Kd for unlabeled heparin, as calculated from competition experiments, was 0.23 microM. Much lower affinity was calculated for unlabeled low molecular weight heparin fragments CY 222 (Kd = 4.3 microM) from competition experiments with [125I]CY 222. The binding reversibility was only partial for unfractionated heparin. Even by chasing with unlabeled compound, a fraction of 25-30% was not dissociable from endothelial cells. This fraction was much lower if incubation was carried out at 4 degrees C. The addition of basic proteins (histones) to the incubation medium greatly enhanced the undissociable binding at 37 degrees C, but not at 4 degrees C. The undissociable fraction of heparin was not available to degradation by purified microbial heparinase. These results suggest that a fraction of bound heparin is internalized by the vascular endothelium.  相似文献   

12.
We have previously reported that exposing cultured Madin Darby canine kidney (MDCK) cells to the polycation protamine (PRO) results in increased short-circuit current and decreased barrier integrity as measured by mannitol permeability and transepithelial electrical resistance. To further investigate the interaction of PRO with the surface of epithelial cells, we labeled PRO with [14C] with use of reductive alkylation. [14C]PRO bound to the cells in a biphasic pattern. Approximately 10% of the [14C]PRO was bound to the cells in the first 5 min, followed by an additional 10% that was bound over the next 25 min. No additional [14C]PRO bound to the cells after the initial 30 min. Binding of [14C]PRO was inhibited by "cold" PRO, which suggested specificity. Binding was also inhibited by polyanions, serum, and albumin, agents previously found to protect MDCK cells from PRO-induced injury. The binding of PRO to MDCK cells was not inhibited by incubation of the MDCK cells with neuraminidase, to remove surface sialic acid residues, or with heparinase, to remove surface heparan sulfate, even though metabolic labeling experiments demonstrated that neuraminidase decreased cell sialic acid and heparinase decreased cell heparan sulfate. Neuraminidase and heparinase offered no protection from PRO injury and had no effect themselves on mannitol permeability. Incubation of the cells with trypsin, however, blunted both the binding of PRO to the cells and the increase in mannitol permeability after exposure of the cells to PRO.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Previous studies have suggested that the attachment of bovine herpesvirus 1 (BHV-1) to permissive cells is mediated by its major glycoproteins B (gB), C (gC), and D (gD). In order to gain further insight into the mechanism of the BHV-1 attachment process, we purified authentic gB, gC, and gD from BHV-1-infected cells and membrane anchor-truncated, soluble gB, gC, and gD from stably transfected cell lines by affinity chromatography and examined their cell-binding properties on Madin-Darby bovine kidney cells. All of the glycoproteins tested exhibited saturable binding to Madin-Darby bovine kidney cells. All of the glycoproteins tested exhibited saturable binding to Madin-Darby bovine kidney cells. Addition of exogenous heparin or treatment of cells with heparinase to remove cellular heparan sulfate (HS) prevented both gC and gB from binding to cells but had no effect on gD binding. An assessment of competition between gB, gC, and gD for cell binding revealed that gC was able to inhibit gB binding, whereas other combinations showed no effect. Cell-bound gC could be dissociated by heparin or heparinase treatment. The response of bound gB to heparin and heparinase treatments differed for the authentic and soluble forms; while soluble gB was susceptible to the treatment, a significant portion of cell-bound authentic gB was resistant to the treatment. Binding affinity analysis showed that soluble gB and both forms of gC and gD each had single binding kinetics with comparable dissociation constants (Kds), ranging from 1.5 x 10(-7) to 5.1 x 10(-7) M, whereas authentic gB exhibited dual binding kinetics with Kd1 = 5.2 x 10(-7) M and Kd2 = 4.1 x 10(-9) M. These results demonstrate that BHV-1 gC binds only to cellular HS, gD binds to a non-HS component, and gB initially binds to HS and then binds with high affinity to a non-HS receptor. Furthermore, we found that while authentic gB was able to inhibit viral plaque formation, soluble gB, which retains the HS-binding property but lacks the high-affinity binding property, was defective in this respect. These results suggest that the interaction between gB and its high-affinity receptor may play a critical role in the virus entry process.  相似文献   

14.
The binding of Apolipoprotein E supplemented triglyceride emulsions to sulfated glycosaminoglycans demonstrated specificity for the carbohydrate polymers. Glucosamine containing glycosaminoglycans with relatively less sulfate had little affinity for the Apo E emulsion whereas those with more sulfate (i.e. heparin and sulfated heparans) effectively bound the emulsion. Galactosamine containing glycosaminoglycans (chondroitin 4 sulfate and dermatan sulfate) demonstrated no binding. The Apo E induced uptake of triglyceride emulsions by hepatocytes was inhibited by highly sulfated polysaccharides (i.e. heparin, dextran sulfate) but other glycosaminoglycans which did not bind the emulsion were ineffective in this inhibition. The same sulfated compounds which inhibited the hepatocyte Apo E emulsion interaction effectively released hepatic lipase from isolated heptic perfusions. Glycosaminoglycan sulfates which did not bind the Apo E supplemented emulsions and did not inhibit hepatocyte association were ineffective in releasing lipase. A heparan mixture isolated from human liver was much less effective in inhibiting Apo E induced association of emulsions with hepatocytes, than heparin. A highly sulfated octasaccharide fraction isolated from bovine liver heparin inhibited more effectively than the human heparans but less than the heparin. Inhibition of Apo E mediated hepatocyte emulsion association was produced by a one hour exposure of the cells to either heparinase or heparanase. The heparanase was more active than the heparinase and both were effective in the presence of protease inhibitors. Enzymes hydrolyzing chondroitin sulfates and hyaluronic acid were ineffective in inhibiting the Apo E induced association. The specific binding of human low density lipoprotein to the hepatocyte was much less effected by the heparanase exposure than the Apo E mediated binding.  相似文献   

15.
Mouse mammary epithelial cells (NMuMG cells) deposit at their basal surfaces an extracellular heparan sulfate-rich proteoglycan that binds to type I collagen. The binding of the purified proteoglycan to collagen was studied by (i) a solid phase assay, (ii) a suspension assay using preformed collagen fibrils, and (iii) a collagen fibril affinity column. The binding interaction occurs at physiological pH and ionic strength and can be inhibited only by salt concentrations that greatly exceed those found physiologically. Binding requires the intact proteoglycan since the protein-free glycosaminoglycan chains will not bind under the conditions of these assays. However, binding is mediated through the heparan sulfate chains as it can be inhibited by block-sulfated polysaccharides, including heparin. Binding requires native collagen structure which may be optimal when the collagen is in a fibrillar configuration. Binding sites on collagen fibrils are saturable, high affinity (Kd approximately 10(-10) M), and selective for heparin-like glycosaminoglycans. Because a culture substratum of type I collagen fibrils causes NMuMG cells to accumulate heparan sulfate proteoglycan into a basal lamina-like layer, binding of heparan sulfate proteoglycans to type I collagen may lead to the formation of a basal lamina and may link the basal lamina to the connective tissue matrix, an association found in basement membranes.  相似文献   

16.
A heparan sulfate glycosaminoglycan chain, biotinylated at its reducing-end, was bound to a streptavidin-coated biochip. Surface plasmon resonance spectroscopy showed a low affinity interaction with antithrombin III (ATIII) when it was flowed over a surface containing heparan sulfate. ATIII bound tightly with high affinity when the same surface was enzymatically modified to using 3-O-sulfotransferase isoform 1 (3-OST-1) in the presence of 3'-phosphoadenosine 5'-phosphosulfate (PAPS). The 3-OST-1 enzyme is involved in heparan sulfate biosynthesis and introduces a critical 3-O-sulfo group into this glycosaminoglycan affording the appropriate pentasaccharide sequence capable of high affinity binding to ATIII. This experiment demonstrates the specific structural modification of a glycosaminoglycan bound to a biochip using a biosynthetic enzyme, suggesting a new approach to rapid screening glycosaminoglycan-protein interactions.  相似文献   

17.
Heparan sulfate proteoglycans (HSPG) are obligatory for receptor binding and mitogenic activity of basic fibroblast growth factor (bFGF). Mutant Chinese hamster ovary cells (pgsA-745) deficient in xylosyltransferase are unable to initiate glycosaminoglycan synthesis and hence can not bind bFGF to low- and high-affinity cell surface receptors. Exposure of pgsA-745 cells to β-D-xylopyranosides containing hydrophobic aglycones resulted in restoration of bFGF binding in a manner similar to that induced by soluble heparin or by heparan sulfate (HS) normally associated with cell sulfate. Restoration of bFGF binding correlated with the ability of the β-D-xylosides to prime the synthesis of heparan sulfate. Thus, both heparan sulfate synthesis and bFGF receptor binding were induced by low concentrations (10–30 μM) of estradiol-β-D-xyloside and naphthyl-β-D-xyloside, but not by cis/trans-decahydro-2-naphthyl-β-D-xyloside, which at low concentration primes mainly chondroitin sulfate. The obligatory involvement of xyloside-primed heparan sulfate in restoration of bFGF-receptor binding was also demonstrated by its sensitivity to heparinase treatment and by the lack of restoration activity in CHO cell mutants that lack enzymatic activities required to form the repeating disaccharide unit characteristic of heparan sulfate. Xyloside-primed heparan sulfate binds to the cell surface. Restoration of bFGF receptor binding was induced by both soluble and cell bound xyloside-primed heparan sulfate and was abolished in cells that were exposed to 0.5–1.0 M NaCl prior to the bFGF binding reaction. These results indicate that heparan sulfate chains produced on xyloside primers behave like heparan sulfate chains attached to cellular core proteins in terms of affinity for bFGF and ability to function as low-affinity sites in a dual receptor mechanism characteristic of bFGF and other heparin-binding growth promoting factors.  相似文献   

18.
The purification of two heparitinases and a heparinase, in high yields from Flavobacterium heparinum was achieved by a combination of molecular sieving and cation-exchange chromatography. Heparinase acts upon N-sulfated glucosaminido-L-iduronic acid linkages of heparin. Substitution of N-sulfate by N-acetyl groups renders the heparin molecule resistant to degradation by the enzyme. Heparitinase I acts on N-acetylated or N-sulfated glucosaminido-glucuronic acid linkages of the heparan sulfate. Sulfate groups at the 6-position of the glucosamine moiety of the heparan sulfate chains seem to be impeditive for heparitinase I action. Heparitinase II acts upon heparan sulfate producing disulfated, N-sulfated and N-acetylated-6-sulfated disaccharides, and small amounts of N-acetylated disaccharide. These and other results suggest that heparitinase II acts preferentially upon N,6-sulfated glucosaminido-glucuronic acid linkages. The total degradation of heparan sulfate is only achieved by the combined action of both heparitinases. The 13C NMR spectra of the disaccharides formed from heparan sulfate and a heparin oligosaccharide formed by the action of the heparitinases are in accordance to the proposed mode of action of the enzymes. Comparative studies of the enzymes with the commercially available heparinase and heparitinase are described.  相似文献   

19.
The specificity, affinity and stoichiometry of the interaction between avidin and glycosaminoglycans (GAGs) have been investigated using heparin-coated microtiter-plate assays, a filter binding assay and surface plasmon resonance (SPR) analysis using a BIAcore 2000 biosensor. Avidin binds heparin and heparan sulfate, and chondroitin-4-sulfate, chondroitin-6-sulfate, dermatan sulfate or hyaluronan were unable to compete for binding. Highest-affinity binding was observed with heparin, and weaker binding was seen when using heparan sulfate or low molecular weight heparin preparations. This indicated that only specific polysaccharide structures tightly interact with avidin. Approximately two avidin molecules bind to each heparin molecule with an overall affinity of 160 nM. The interaction is pH dependent, increasing five-fold upon decreasing the pH from 7.5 to 5.5, while binding was negligible at pH 9. We demonstrate the potential of fluorescent avidin derivatives as a tool for the detection of heparin and heparan sulfates on surfaces by application to both heparin immobilized on polystyrene plates and heparan sulfate on cell surfaces.  相似文献   

20.
Glycosaminoglycan-binding proteins, with specific emphasis on dermatan sulfate, have been investigated in human plasma by affinity chromatography, mass spectrometry and Western blotting. Diluted plasma was applied to affinity columns and bound protein was eluted with 500 mM NaCl. Dermatan sulfate and heparan sulfate bound 7% of the total protein. Heparin bound 22% of the total protein, but chondroitin sulfate A bound only 0.23%. Mass spectrometric analysis identified 20 proteins as dermatan-sulfate-binding proteins, most of which were confirmed by Western blotting. Some of these binding proteins, such as fibrinogen, fibronectin, apolipoprotein B, LMW kininogen, inter-alpha-trypsin inhibitor, and factor H, were degraded to various extents during the chromatography step, but this degradation could be prevented by the inclusion of a serine protease inhibitor. The protein fraction binding to the dermatan sulfate column showed amidase activity, whereas that binding to the heparan sulfate and heparin columns showed 1/2 and 1/20, respectively, of the activity of the dermatan sulfate binding fraction. Dermatan sulfate was similar to heparan sulfate with respect to its capacity to bind plasma proteins and its activation of protease, but differed from chondroitin sulfate and heparin in these properties.  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