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 26 毫秒
1.
Heparan sulfates from Swiss mouse 3T3 and SV3T3 cells: O-sulfate difference   总被引:7,自引:0,他引:7  
K L Keller  J M Keller  J N Moy 《Biochemistry》1980,19(11):2529-2536
A difference in the extent of sulfation between the heparan sulfate isolated from Swiss 3T3 mouse cells and that from Swiss 3T3 cells transformed by the DNA virus SV40 has been reported previously. This variance is manifested by different chromatographic and electrophoretic properties. Heparan sulfates from the two cell types were treated with nitrous acid under conditions that gave selective deaminative cleavage of glucosaminyl residues with sulfated amino groups in order to define the nature of the difference in sulfation further. The O-sulfate containing fragments from the heparan sulfates were compared by gel filtration and ion-exchange chromatography. The results showed that the 3T3 heparan sulfate contains 8% more O-sulfate than does the SV3T3 heparan sulfate. Analysis of uronic acids revealed that both types of heparan sulfates contain 45% L-iduronic acid and 55% D-glucuronic acid. These and other observations indicate that the primary difference in sulfation between the 3T3 and SV3T3 heparan sulfates lies in the extent of O-sulfation.  相似文献   

2.
Safaiyan F  Lindahl U  Salmivirta M 《Biochemistry》2000,39(35):10823-10830
The N-sulfated regions (NS domains) represent the modified sequences of heparan sulfate chains and mediate interactions of the polysaccharide with proteins. We have investigated the relationship between the type/extent of polymer modification and the length of NS domains in heparan sulfate species from human aorta, bovine kidney, and cultured NMuMG and MDCK cells. C5 epimerization of D-glucuronic acid to L-iduronic acid was found to be extensive and essentially similar in all heparan sulfate species studied, regardless of domain size, whereas the subsequent 2-O-sulfation of the formed iduronic acid residues varies appreciably. In aorta heparan sulfate, up to 90% of the formed iduronate residues were 2-O-sulfated, whereas in kidney heparan sulfate 2-O-sulfation occurred only in 相似文献   

3.
Heparan sulfate is a highly sulfated polysaccharide abundantly present in the extracellular matrix. Heparan sulfate consists of a disaccharide repeating unit of glucosamine and glucuronic and iduronic acid residues. The functions of heparan sulfate are largely dictated by its size as well as the sulfation patterns. Heparanase is an enzyme that cleaves heparan sulfate polysaccharide into smaller fragments, regulating the functions of heparan sulfate. Understanding the substrate specificity plays a critical role in dissecting the biological functions of heparanase and heparan sulfate. The prevailing view is that heparanase recognizes specific sulfation patterns in heparan sulfate. However, emerging evidence suggests that heparanase is capable of varying its substrate specificities depending on the saccharide structures around the cleavage site. The plastic substrate specificity suggests a complex role of heparanase in regulating the structures of heparan sulfate in matrix biology.  相似文献   

4.
Heparan sulfate found in the cerebral plaques of Alzheimer's disease binds to beta-amyloid (Abeta) fibrils. This interaction has been proposed to enhance fibril deposition and mediate Abeta-induced glia activation and neurotoxicity. On the other hand, heparan sulfate augments signaling of fibroblast growth factor-2 (FGF-2), a neuroprotective factor that antagonizes the neurotoxic effects of Abeta. We defined structures in heparan sulfate from human cerebral cortex that bind Abeta fibrils. The minimal binding site is found in N-sulfated hexasaccharide domains and contains critical 2-O-sulfated iduronic acid residues. By contrast, binding of Abeta monomers requires, in addition, 6-O-sulfate groups on glucosamine residues. The binding specificity of fibrillar Abeta is shared by FGF-2, and we here show that cerebral heparan sulfate domains selected for binding to Abeta-(1-40) fibrils bind also to FGF-2. These data suggest that neurotoxic and neuroprotective signals may converge by competing for the same binding sites on the heparan sulfate chain.  相似文献   

5.
Heparan sulfate N-sulfotransferase catalyzes the transfer of sulfate groups from adenosine 3'-phosphate, 5'-phosphosulfate to the free amino groups of glucosamine residues in heparan sulfate. We have identified a Chinese hamster ovary cell mutant, designated pgsE-606, which is 3-5-fold defective in N-sulfotransferase activity. The residual enzyme activity is indistinguishable from the wild-type enzyme with respect to Km values for adenosine 3'-phosphate,5'-phosphosulfate and N-desulfoheparin, pH dependence, Arrhenius activation energy, and thermal lability. The mutation is recessive, and mixing experiments indicate that the mutant does not produce soluble antagonists of N-sulfotransferase. Inspection of the heparan sulfate chains from the mutant showed that the extent of N-sulfation is reduced about 2-3-fold. The addition of sulfate to hydroxyl groups on the chain is reduced to a similar extent, suggesting that N-sulfation and O-sulfation are normally coupled. Nitrous acid fragmentation of the chains showed that N-sulfated glucosamine residues are spaced much less frequently than in heparan sulfate from wild-type cells. The close correlation of enzyme activity to the number and position of N-sulfate groups indicates that N-sulfotransferase plays a pivotal role in determining the extent of sulfation of heparan sulfate.  相似文献   

6.
We have examined the size, charge, and sulfation pattern of heparan sulfate in the cell-soluble fraction, membranes, and culture medium of PC12 pheochromocytoma cells cultured in the presence and absence of nerve growth factor (NGF) and compared the structural features of PC12 cell heparan sulfate to that of rat brain at several stages of early postnatal development. Nitrous acid degradation studies revealed significant differences in the distribution of N-sulfate and N-acetyl groups in heparan sulfate present in the PC12 cell-soluble fraction, membranes, and medium and demonstrated that NGF treatment led to an increased proportion of N-sulfated segments in the cell-associated heparan sulfate, although no such change was seen in that released into the culture medium. There was very little change in the N-sulfation of brain heparan sulfate during the first 30 days after birth. In brain, most of the heparan sulfate glucosamine residues are N-sulfated and yield predominantly di- and tetrasaccharide nitrous acid degradation products, whereas PC12 cell heparan sulfate contains large blocks of N-acetylglucosamine residues. There was very little difference in the overall charge or size (approximately 15,000 Da) of heparan sulfate chains between the different PC12 cell fractions or brain, although NGF treatment led to a decrease in the proportion of less-charged chains in the PC12 cell membranes and a small increase in molecular size. Our studies therefore demonstrate the presence in PC12 cells of several pools of heparan sulfate having different structural properties, and that significant alterations in the charge, size, and sulfation pattern of PC12 cell heparan sulfate accompany NGF-induced differentiation and neurite outgrowth.  相似文献   

7.
8.
Cell surface heparan sulfate (HS) proteoglycans are required in development and postnatal repair. Important classes of ligands for HS include growth factors and extracellular matrix macromolecules. For example, the focal adhesion component syndecan-4 interacts with the III(12-14) region of fibronectin (HepII domain) through its HS chains. The fine structure of HS is critical to growth factor responses, and whether this extends to matrix ligands is unknown but is suggested from in vitro experiments. Cell attachment to HepII showed that heparin oligosaccharides of >or=14 sugar residues were required for optimal inhibition. The presence of N-sulfated glucosamine in the HS was essential, whereas 2-O-sulfation of uronic acid or 6-O-sulfation of glucosamine had marginal effects. In the more complex response of focal adhesion formation through syndecan-4, N-sulfates were again required and also glucosamine 6-O-sulfate. The significance of polymer N-sulfation and sulfated domains in HS was confirmed by studies with mutant Chinese hamster ovary cells where heparan sulfation was compromised. Finally, focal adhesion formation was absent in fibroblasts synthesizing short HS chains resulting from a gene trap mutation in one of the two major glucosaminoglycan polymerases (EXT1). Several separate, specific properties of cell surface HS are therefore required in cell adhesion responses to the fibronectin HepII domain.  相似文献   

9.
The medium and cell surface heparan sulfates isolated from SV40-transformed Swiss mouse 3T3 cells were examined in the presence and absence of 1.0 mM p-nitrophenyl-beta-D-xyloside. Incubation of the SV3T3 cells with this beta-xyloside resulted in: (a) a 4- to 5-fold reduction in the molecular weight distribution of medium heparan sulfate, (b) a 10-fold increase in the total synthesis of medium heparan sulfate, and (c) a small reduction in cell growth. There was little, if any, change in either the total level of synthesis or the molecular weight distribution of cell surface heparan sulfate. The covalent association of the beta-xyloside to the medium heparan sulfate was demonstrated by an analysis of the medium heparan sulfate produced by cells grown in the presence of [35S]sulfate and the fluorogenic beta-xyloside, 4-methylumbelliferyl-beta-D-xyloside. Treatment of the purified radiolabeled and fluorogenic heparan sulfate with either nitrous acid or heparitinase resulted in a decrease in the molecular weight of both radiolabeled and fluorogenic material. The data presented in this paper are discussed with respect to both the structure of heparan sulfate and the putative role of heparan sulfate in cell social behavior.  相似文献   

10.
Heparan sulfate polymerization and modification take place in the Golgi compartment. The modification reactions are initiated by glucosaminyl N-deacetylase/N-sulfotransferase (NDST), a bifunctional enzyme that removes N-acetyl groups from selected N-acetyl-d-glucosamine units followed by N-sulfation of the generated free amino groups. Four isoforms of NDST have been identified. NDST-1 and -2 have a wide and largely overlapping tissue distribution, but it is not known if they can act on the same heparan sulfate chain. We have introduced point mutations into NDST-1 cDNA, which selectively destroy the N-deacetylase or N-sulfotransferase activity of the enzyme [Wei, Z., and Swiedler, S. J. (1999) J. Biol. Chem. 274, 1966-70 and Sueyoshi, T., et al. (1998) FEBS Lett. 433, 211-4]. Stable 293 cell lines expressing the NDST-1 mutants were then generated. Structural analyses of heparan sulfate synthesized by these cells and by cells overexpressing wild-type NDST-1 demonstrate that the N-deacetylation step is not only prerequisite but also rate-limiting, determining the degree of N-sulfation. Transfection of mutant NDST-1 lacking N-deacetylase activity had no effect on heparan sulfate sulfation, while cells expressing wild-type enzyme or NDST-1 lacking N-sulfotransferase activity both resulted in the production of oversulfated heparan sulfate. Since no increase in the amount of N-unsubstituted glucosamine residues was seen after transfection of the mutant lacking N-sulfotransferase activity, the results also suggest that two different enzyme molecules can act on the same glucosamine unit. In addition, we show that oversulfation of heparan sulfate produced by cells tranfected with wild-type NDST-1 or the mutant lacking N-sulfotranferase activity results in decreased sulfation of chondroitin sulfate.  相似文献   

11.
Heparan sulfate mediates numerous complex biological processes. Its action critically depends on the amount and the positions of O-sulfate groups (iduronyl 2-O-sulfates, glucosaminyl 6-O- and 3-O-sulfates) that form binding sites for proteins. The structures and distribution of these protein-binding domains are influenced by the expression and substrate specificity of heparan sulfate biosynthetic enzymes. We describe a general approach to assess substrate specificities of enzymes involved in glycosaminoglycan metabolism, here applied to 6-O-sulfotransferases involved in heparan sulfate biosynthesis. To understand how 2-O-sulfation affects subsequent 6-O-sulfation reactions, the substrate specificity of 6-O-sulfotransferase 3 was probed using substrates from a heparin-based octasaccharide library. Purified 3H-labeled N-sulfated octasaccharides from a library designed to sample 2-O-sulfated motifs were used as sulfate acceptors, 3'-phosphoadenosine 5'-phosphosulfate as sulfate donor, and cell extract from 6-O-sulfotransferase 3-overexpressing 293 cells as enzyme source in the 6-O-sulfotransferase-catalyzed reactions. The first 6-O-sulfate group was preferentially incorporated at the internal glucosamine unit of the octasaccharide substrate. As the reaction proceeded, the octasaccharides acquired three 6-O-sulfate groups. The specificities toward competing octasaccharide substrates, for 6-O-sulfotransferase 2 and 6-O-sulfotransferase 3, were determined using overexpressing 293 cell extracts and purified octasaccharides. Both 6-O-sulfotransferases showed a preference for 2-O-sulfated substrates. The specificity toward substrates with two to three 2-O-sulfate groups was three to five times higher as compared with octasaccharides with no or one 2-O-sulfate group.  相似文献   

12.
Incubation of microsomal fractions with labelled 3'-phosphoadenylyl sulfate results in incorporation of [35S]sulfate into endogenous glycosaminoglycans. Specific radioactivity observed incorporated into heparan sulfate chains is 10-fold greater than that incorporated into chondro?tin sulfate chains. This is in agreement with the results obtained for glycosylation of glycosaminoglycans in arterial wall membrane fractions. Sulfation of heparan sulfate was studied since it contains N- and O-sulfate groups in contrast with the other sulfated glycosaminoglycans which contain only O-sulfate groups. Sulfation of heparan sulfate occurs rapidly, since sulfate incorporation is detected after exposure for only 0.5 min. Heparan sulfate was identified on the basis of its resistance to hyaluronidase and chondro?tin ABC lyase, its susceptibility to heparitinase, its sensitivity to nitrous acid and the presence of glucosamine as the only hexosamine. The chemical composition of the purified heparan sulfate fractions provides evidence for the high degree of sulfation of its chains. Studies into the distribution of sulfate residues on heparan sulfate at different times of sulfation indicate that N-sulfate groups are not randomly introduced into the polymer. The relationship between the processes of N- and O-sulfation was studied. The present results demonstrate that preferential N-sulfation is obtained for incorporation of labelled precursor over a short period, the O-sulfation occurring on previously N-sulfated heparan sulfate.  相似文献   

13.
Sulf1 and Sulf2 are two heparan sulfate 6-O-endosulfatases that regulate the activity of multiple growth factors, such as fibroblast growth factor and Wnt, and are essential for mammalian development and survival. In this study, the mammalian Sulfs were functionally characterized using overexpressing cell lines, in vitro enzyme assays, and in vivo Sulf knock-out cell models. Analysis of subcellular Sulf localization revealed significant differences in enzyme secretion and detergent solubility between the human isoforms and their previously characterized quail orthologs. Further, the activity of the Sulfs toward their native heparan sulfate substrates was determined in vitro, demonstrating restricted specificity for S-domain-associated 6S disaccharides and an inability to modify transition zone-associated UA-GlcNAc(6S). Analysis of heparan sulfate composition from different cell surface, shed, glycosylphosphatidylinositol-anchored and extracellular matrix proteoglycan fractions of Sulf knock-out cell lines established differential effects of Sulf1 and/or Sulf2 loss on nonsubstrate N-, 2-O-, and 6-O-sulfate groups. These findings indicate a dynamic influence of Sulf deficiency on the HS biosynthetic machinery. Real time PCR analysis substantiated differential expression of the Hs2st and Hs6st heparan sulfate sulfotransferase enzymes in the Sulf knock-out cell lines. Functionally, the changes in heparan sulfate sulfation resulting from Sulf loss were shown to elicit significant effects on fibroblast growth factor signaling. Taken together, this study implicates that the Sulfs are involved in a potential cellular feed-back mechanism, in which they edit the sulfation of multiple heparan sulfate proteoglycans, thereby regulating cellular signaling and modulating the expression of heparan sulfate biosynthetic enzymes.  相似文献   

14.
Heparan sulfate from the surface of a variety of mouse cells at different cell densities was examined by ion-exchange chromatography. The results of this analysis show that: (1) The heparan sulfate from new isolates of Swiss 3T3 cells transformed by SV40 virus (a DNA tumor virus) elutes from DEAE-cellulose at a lower ionic strength than that from the parent cell type. This finding confirms our earlier observation with an established SV40-transformed cell line (Underhill and Keller, '75) and eliminates the possibility that this change is caused by extended passage in culture. (2) For both parent and transformed 3T3 cells, the heparan sulfates from low and high density cultures were the same as judged by chromatography on DEAE-cellulose. This result demonstrates that the transformation-dependent change which we have observed is independent of cell density. (3) The heparan sulfate from Balb/c 3T3 cells transformed with Kirsten murine sarcoma virus (an RNA tumor virus) elutes from DEAE-cellulose prior to that from parent Balb/c 3T3 cells. This result extends the transformation dependent change in heparan sulfate to the Balb/c 3T3 cell line and to cells transformed with an RNA virus.  相似文献   

15.
We show here that the endothelial cell-line ECV 304 expresses the heparan sulfate proteoglycan glypican-1. The predominant cellular glycoform carries truncated side-chains and is accompanied by heparan sulfate oligosaccharides. Treatment with brefeldin A results in accumulation of a glypican proteoglycan with full-size side-chains while the oligosaccharides disappear. During chase the glypican proteoglycan is converted to partially degraded heparan sulfate chains and chain-truncated proteoglycan, both of which can be captured by treatment with suramin. The heparan sulfate chains in the intact proteoglycan can be depolymerized by nitrite-dependent cleavage at internally located N-unsubstituted glucosamine moieties. Inhibition of NO-synthase or nitrite-deprivation prevents regeneration of intact proteoglycan from truncated precursors as well as formation of oligosaccharides. In nitrite-deprived cells, formation of glypican proteoglycan is restored when NO-donor is supplied. We propose that, in recycling glypican-1, heparan sulfate chains are cleaved at or near glucosamines with unsubstituted amino groups. NO-derived nitrite is then required for the removal of short, nonreducing terminal saccharides containing these N-unsubstituted glucosamine residues from the core protein stubs, facilitating re-synthesis of heparan sulfate chains.  相似文献   

16.
IntroductionSulf1 and Sulf2 are cell surface sulfatases, which remove specific 6-O-sulfate groups from heparan sulfate (HS) proteoglycans, resulting in modulation of various HS-dependent signaling pathways. Both Sulf1 and Sulf2 knockout mice show impairments in brain development and neurite outgrowth deficits in neurons.ConclusionSulfs introduce dynamic changes in HS proteoglycan sulfation patterns of the postnatal cerebellum, thereby orchestrating fundamental mechanisms underlying brain development.  相似文献   

17.
3-O-Sulfation of glucosamine by heparan sulfate D-glucosaminyl 3-O-sulfotransferase (3-OST-1) is the key modification in anticoagulant heparan sulfate synthesis. However, the heparan sulfates modified by 3-OST-2 and 3-OST-3A, isoforms of 3-OST-1, do not have anticoagulant activity, although these isoforms transfer sulfate to the 3-OH position of glucosamine residues. In this study, we characterize the substrate specificity of purified 3-OST-3A at the tetrasaccharide level. The 3-OST-3A enzyme was purified from Sf9 cells infected with recombinant baculovirus containing 3-OST-3A cDNA. Two 3-OST-3A-modified tetrasaccharides were purified from the 3-O-(35)S-sulfated heparan sulfate that was digested by heparin lyases. These tetrasaccharides were analyzed using nitrous acid and enzymatic degradation combined with matrix-assisted laser desorption/ionization-mass spectrometry. Two novel tetrasaccharides were discovered with proposed structures of DeltaUA2S-GlcNS-IdoUA2S-[(35)S]GlcNH(2)3S and DeltaUA2S-GlcNS-IdoUA2S-[3-(35)S]GlcNH(2)3S6S . The results demonstrate that 3-OST-3A sulfates N-unsubstituted glucosamine residues, and the 3-OST-3A modification sites are probably located in defined oligosaccharide sequences. Our study suggests that oligosaccharides with N-unsubstituted glucosamine are precursors for sulfation by 3-OST-3A. The intriguing linkage between N-unsubstituted glucosamine and the 3-O-sulfation by 3-OST-3A may provide a clue to the potential biological functions of 3-OST-3A-modified heparan sulfate.  相似文献   

18.
Three major pools of heparan sulfate have been isolated from cultures of Swiss mouse 3T3 and SV40-transformed 3T3 cells: cell-surface, medium, and intracellular heparan sulfates. The cell-surface heparan sulfate is a high molecular weight proteogylcan which is partially degraded by pronase. Before pronase treatment, it has a peak molecular weight (as estimated by gel filtration) of appox. 7.2 · 105 in contrast to only 2.4 · 105 after pronase treatment. The medium heparan sulfate appears to be similar in structure to the cell-surface heparan sulfate, since they coelute on Bio-Gel A-15m and DEAE-cellulose, and are both proteoglycans. In contrast, the intracellular heparan sulfate has a low molecular weight (6.0 · 103) and has little if any attached protein. Both the medium and intracellular heparan sulfate exhibit the transformation-associated change in structure reported earlier for cell-surface heparan sulfate (Underhill, C.B. and Keller, J.M. (1975) Biochem. Biophys. Res. Commun. 63, 448–454). This transformation-associated change, detected by DEAE-cellulose chromatography is not the result of changes in either molecular weight or protein core. Cellulose acetate electrophoresis of the cell-surface heparan sulfate at pH 1 suggests that the transformation-associated change in structure is due to a difference in sulfate content. Both types of heparan sulfate are produced in mixed cultures ot 3T3 and SV3T3 cells, indicating that neither serum factors in the culture medium nor secreted cell products are responsible for the transformation-associated change in heparan sulfate structure. The presented date are discussed with respect to the postulated role of heparan sulfate in cell social behavior.  相似文献   

19.
Three major pools of heparan sulfate have been isolated from cultures of Swiss mouse 3T3 and SV40-transformed 3T3 cells: cell-surface, medium, and intracellular heparan sulfates. The cell-surface heparan sulfate is a high molecular weight proteoglycan which is partially degraded by pronase. Before pronase treatment, it has a peak molecular weight (as estimated by gel filtration) of approx. 7.2 . 10(5) in contrast to only 2.4 . 10(5) after pronase treatment. The medium heparan sulfate appears to be similar in structure to the cell-surface heparan sulfate, since they coelute on Bio-Gel A-15m and DEAE-cellulose, and are both proteoglycans. In contrast, the intracellular heparan sulfate has a low molecular weight (6.0 . 10(3)) and has little if any attached protein. Both the medium and intracellular heparan sulfate exhibit the transformation-associated change in structure reported earlier for cell-surface heparan sulfate (Underhill, C.B. and Keller, J.M. )1975) Biochem. Biophys. Res. Commun. 63, 448--454). This transformation-associated change, detected by DEAE-cellulose chromatography is not the result of changes in either molecular weight or protein core. Cellulose acetate electrophoresis of the cell-surface heparan sulfate at pH 1 suggests that the transformation-associated change in structure is due to a difference in sulfate content. Both types of heparan sulfate are produced in mixed cultures of 3T3 and SV3T3 cells, indicating that neither serum factors in the culture medium nor secreted cell products are responsible for the transformation-associated change in heparan sulfate structure. The presented data are discussed with respect to the postulated role of heparan sulfate in cell social behavior.  相似文献   

20.
Monensin is a monovalent metal ionophore that affects the intracellular translocation of secretory proteins at the level of trans-Golgi cisternae. Exposure of endothelial cells to monensin results in the synthesis of heparan sulfate and chondroitin sulfate with a lower degree of sulfation. The inhibition is dose dependent and affects the ratio [35S]-sulfate/[3H]-hexosamine of heparan sulfate from both cells and medium, with no changes in their molecular weight. By the use of several degradative enzymes (heparitinases, glycuronidase, and sulfatases) the fine structure of the heparan sulfate synthesized by control and monensin-treated cells was investigated. The results have shown that among the six heparan sulfate disaccharides there is a specific decrease of the ones bearing a sulfate ester at the 6-position of the glucosamine moiety. All other biosynthetic steps were not affected by monensin. The results are indicative that monensin affects the hexosamine C-6 sulfation, and that this sterification is the last step of the heparan sulfate biosynthesis and should occur at the trans-Golgi compartment.  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