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 875 毫秒
1.
The rare N-unsubstituted glucosamine (GlcNH (3)(+)) residues in heparan sulfate (HS) have important biological and pathophysiological roles. However, it is difficult to prepare naturally-occuring, GlcNH(3)(+)-containing oligosaccharides from HS because of their low abundance, as well as the inherent problems in both excising and identifying them. Therefore, the ability to chemically generate a series of structurally-defined oligosaccharides containing GlcNH(3)(+) residues would greatly contribute to investigating their natural role in HS. In this study, a series of heparin/HS oligosaccharides, from dp6 up to dp16 in length that possess internal GlcNH(3)(+) residues were prepared by a combination of chemical modification and heparinase I digestion. Purification and structural analysis of the major species derived from the octa- to dodeca-saccharide size fractions indicated the introduction of between 1 and 3 internal GlcNH(3)(+) residues per oligosaccharide. In addition, a GlcNH(3)(+) residue was selectively introduced into an internal position in a tetrasaccharide species by direct chemical modification. This selectivity has potential as an alternative procedure for preparing internally-modified oligosaccharides of various lengths. The utility of such oligosaccharides was demonstrated by a comparison of the binding of three different tetrasaccharide species containing 0, 1 and 2 free amino groups to the NK1 truncated variant of hepatocyte growth factor/scatter factor.  相似文献   

2.
Upon induction with heparin, Flavobacterium heparinum synthesizes and secretes into its periplasmic space heparinase I (EC 4.2.2.7), heparinase II, and heparinase III (heparitinase; EC 4.2.2.8). Heparinase I degrades heparin, and heparinase II degrades both heparin and heparan sulfate, while heparinase III degrades heparan sulfate predominantly. We isolated the genes encoding heparinases II and III (designated hepB and hepC, respectively). These genes are not contiguous with each other or with the heparinase I gene (designated hepA). hepB and hepC were found to contain open reading frames of 2,316 and 1,980 bp, respectively. Enzymatic removal of pyroglutamate groups permitted sequence analysis of the amino termini of both mature proteins. It was determined that the mature forms of heparinases II and III contain 746 and 635 amino acids, respectively, and have calculated molecular weights of 84,545 and 73,135, respectively. The preproteins have signal sequences consisting of 26 and 25 amino acids. Truncated hepB and hepC genes were used to produce active, mature heparinases II and III in the cytoplasm of Escherichia coli. When these enzymes were expressed at 37 degrees C, most of each recombinant enzyme was insoluble, and most of the heparinase III protein was degraded. When the two enzymes were expressed at 25 degrees C, they were both present predominantly in a soluble, active form.  相似文献   

3.
We compare here the structural and functional properties of heparan sulfate (HS) chains from both male or female adult mouse liver through a combination of molecular sieving, enzymatic cleavage, and strong anion exchange-HPLC. The results demonstrated that male and female HS chains are significantly different by a number of parameters; size determination showed that HS chain lengths were ~100 and ~22 kDa, comprising 30-40 and 6-8 disaccharide repeats, respectively. Enzymatic depolymerization and disaccharide composition analyses also demonstrated significant differences in domain organization and fine structure. N-Unsubstituted glucosamine (ΔHexA-GlcNH(3)(+), ΔHexA-GlcNH(3)(+)(6S), ΔHexA(2S)-GlcNH(3)(+), and N-acetylglucosamine (ΔHexA-GlcNAc) are the predominant disaccharides in male mouse liver HS. However, N-sulfated glucosamine (ΔHexA-GlcNSO(3)) is the predominant disaccharide found in female liver. These structurally different male and female liver HS forms exert differential effects on human mesenchymal cell proliferation and subsequent osteogenic differentiation. The present study demonstrates the potential usefulness of gender-specific liver HS for the manipulation of human mesenchymal cell properties, including expansion, multipotentiality, and subsequent matrix mineralization. Our results suggest that HS chains show both tissue- and gender-specific differences in biochemical composition that directly reflect their biological activity.  相似文献   

4.
Pojasek K  Shriver Z  Hu Y  Sasisekharan R 《Biochemistry》2000,39(14):4012-4019
The heparinases from Flavobacterium heparinum are powerful tools in understanding how heparin-like glycosaminoglycans function biologically. Heparinase III is the unique member of the heparinase family of heparin-degrading lyases that recognizes the ubiquitous cell-surface heparan sulfate proteoglycans as its primary substrate. Given that both heparinase I and heparinase II contain catalytically critical histidines, we examined the role of histidine in heparinase III. Through a series of diethyl pyrocarbonate modification experiments, it was found that surface-exposed histidines are modified in a concentration-dependent fashion and that this modification results in inactivation of the enzyme (k(inact) = 0.20 +/- 0.04 min(-)(1) mM(-)(1)). The DEPC modification was pH dependent and reversible by hydroxylamine, indicating that histidines are the sole residue being modified. As previously observed for heparinases I and II, substrate protection experiments slowed the inactivation kinetics, suggesting that the modified residue(s) was (were) in or proximal to the active site of the enzyme. Proteolytic mapping experiments, taken together with site-directed mutagenesis studies, confirm the chemical modification experiments and point to two histidines, histidine 295 and histidine 510, as being essential for heparinase III enzymatic activity.  相似文献   

5.
The molecular structure of human skin fibroblast heparan sulphate was examined by specific chemical or enzymic depolymerization and high-resolution separation of the resulting oligosaccharides and disaccharides. Important features of the molecular organization, disaccharide composition and O-sulphate disposition of this heparan sulphate were identified. Analysis of the products of HNO2 hydrolysis revealed a polymer in which 53% of disaccharide units were N-acetylated and 47% N-sulphated, with an N-/O-sulphate ratio of 1.8:1. These two types of disaccharide unit were mainly located in separate domains. Heparitinase and heparinase scission indicated that the iduronate residues (37% of total hexuronate) were largely present in contiguous disaccharide sequences of variable size that also contained the majority of the N-sulphate groups. Most of the iduronate residues (approx. 70%) were non-sulphated. About 8-10% of disaccharide units were cleaved by heparinase, but only a minority of these originated from contiguous sequences in the intact polymer. Trisulphated disaccharide units [alpha-N-sulpho-6-sulphoglucosaminyl-(1----4)-iduronate 2-sulphate], which are the major structural units in heparin, made up only 3% of the disaccharide units in heparan sulphate. O-Sulphate groups (approx. 26 per 100 disaccharide units) were distributed almost evenly among C-6 of N-acetylglucosamine, C-2 of iduronate and C-6 of N-sulphated glucosamine residues. The results indicate that the sulphated regions of heparan sulphate have distinctive and potentially variable structural characteristics. The high content of non-sulphated iduronate in this heparan sulphate species suggests a conformational versatility that could have important implications for the biological properties of the polymer.  相似文献   

6.
A novel type of heparinase (heparin lyase, no EC number) has been purified from Bacteroides stercoris HJ-15, isolated from human intestine, which produces three kinds of heparinases. The enzyme was purified to apparent homogeneity by a combination of QAE-cellulose, DEAE-cellulose, CM-Sephadex C-50, hydroxyapatite, and HiTrap SP chromatographies with a final specific activity of 19.5 mmol/min/mg. It showed optimal activity at pH 7.2 and 45 degrees C and the presence of 300 mM KCl greatly enhanced its activity. The purified enzyme activity was inhibited by Cu(2+), Pb(2+), and some agents that modify histidine and cysteine residues, and activated by reducing agents such as dithiothreitol and 2-mercaptoethanol. This purified Bacteroides heparinase is an eliminase that shows its greatest activity on bovine intestinal heparan sulfate, and to a lesser extent on porcine intestinal heparan sulfate and heparin. This enzyme does not act on acharan sulfate but de-O-sulfated acharan sulfate and N-sulfoacharan sulfate were found to be poor substrates. The substrate specificity of this enzyme is similar to that of Flavobacterial heparinase II. However, an internal amino acid sequence of the purified Bacteroides heparinase shows significant (73%) homology to Flavobacterial heparinase III and only 43% homology to Flavobacterial heparinase II. These findings suggest that the Bacteroidal heparinase is a novel enzyme degrading GAGs.  相似文献   

7.
High performance liquid chromatography was performed by an ion-pair reversed-phase method of six standard unsaturated disaccharides derived from heparan sulfate and heparin. Separation of delta Di-GlcNAc, delta Di-GlcN(2S), delta Di-GlcNAc(6S), delta Di-GlcN(2,6- or 2,2'-diS) and delta Di-GlcN(2,6,2'-triS) was achieved on a column of Jasco SC-02 with 10 mM tetrabutylammonium phosphate (pH 7.0) containing 30 or 47% methanol as a mobile phase. delta Di-GlcN(2,6-diS) and delta Di-GlcN(2,2'-diS) were separated on the same column with 35 mM triethylamine phosphate (pH 5.3). Four preparations (BL-1.0-1, BL-1.0-2, BL-1.0-3, and BL-1.25-1) separated from crude bovine lung heparan sulfate, a standard bovine lung heparan sulfate (BL-ST), bovine kidney heparan sulfate 1.0 M Fr and 1.25 M Fr (BK-1.0 and BK-1.25), and porcine kidney heparan sulfate 1.0 M Fr (PK-1.0) were digested with a mixture of heparinase, and heparitinases 1 and 2. The resulting foregoing unsaturated disaccharides in the digests were analyzed by the above HPLC procedures. The proportions of the unsaturated disaccharides in the digests of BL-1.25-1 and BL-ST were similar, but those of the others differed from each other. It is noteworthy that delta Di-GlcNAc plus delta Di-GlcNAc(6S) in the digest of BL-1.0-1 was approximately 95% of the total unsaturated disaccharides. Small amounts of delta Di-GlcN (2,6,2'-triS) were found in all the samples. It was found that delta Di-GlcN(2,2'-diS) was a prominent component in the disulfated unsaturated disaccharides from BL-1.25-1 and BK-1.25.  相似文献   

8.
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   

9.
Functional properties of heparan sulfate (HS) are generally ascribed to the sulfation pattern of the polysaccharide. However, recently reported functional implications of rare N-unsubstituted glucosamine (GlcNH(2)) residues in native HS prompted our structural characterization of sequences around such residues. HS preparations were cleaved with nitrous acid at either N-sulfated or N-unsubstituted glucosamine units followed by reduction with NaB(3)H(4). The labeled products were characterized following complementary deamination steps. The proportion of GlcNH(2) units varied from 0.7-4% of total glucosamine in different HS preparations. The GlcNH(2) units occurred largely clustered at the polysaccharide-protein linkage region in intestinal HS, also more peripherally in aortic HS. They were preferentially located within N-acetylated domains, or in transition sequences between N-acetylated and N-sulfated domains, only 20-30% of the adjacent upstream and downstream disaccharide units being N-sulfated. The nearest downstream (toward the polysaccharide-protein linkage) hexuronic acid was invariably GlcUA, whereas the upstream neighbor could be either GlcUA or IdoUA. The highly sulfated but N-unsubstituted disaccharide unit, -IdoUA2S-GlcNH(2)6S-, was detected in human renal and porcine intestinal HS, but not in HS from human aorta. These results are interpreted in terms of a biosynthetic mechanism, whereby GlcNH(2) residues are formed through regulated, incomplete action of an N-deacetylase/N-sulfotransferase enzyme.  相似文献   

10.
Examination of the substrate specificity of heparin and heparan sulfate lyases   总被引:15,自引:0,他引:15  
We have examined the activities of different preparations of heparin and heparan sulfate lyases from Flavobacterium heparinum. The enzymes were incubated with oligosaccharides of known size and sequence and with complex polysaccharide substrates, and the resulting degradation products were analyzed by strong-anion-exchange high-performance liquid chromatography and by oligosaccharide mapping using gradient polyacrylamide gel electrophoresis. Heparinase (EC 4.2.2.7) purified in our laboratory and a so-called Heparinase I (Hep I) from a commercial source yielded similar oligosaccharide maps with heparin substrates and displayed specificity for di- or trisulfated disaccharides of the structure----4)-alpha-D-GlcNp2S(6R)(1----4)-alpha-L-IdoAp2S( 1----(where R = O-sulfo or OH). Oligosaccharide mapping with two different commercial preparations of heparan sulfate lyase [heparitinase (EC 4.2.2.8)] indicated close similarities in their depolymerization of heparan sulfate. Furthermore, these enzymes only degraded defined oligosaccharides at hexosaminidic linkages with glucuronic acid:----4)-alpha-D-GlcNpR(1----4)-beta-D-GlcAp(1----(where R = N-acetamido or N-sulfo). The enzymes showed activity against solitary glucuronate-containing disaccharides in otherwise highly sulfated domains including the saccharide sequence that contains the antithrombin binding region in heparin. A different commercial enzyme, Heparinase II (Hep II), displayed a broad spectrum of activity against polysaccharide and oligosaccharide substrates, but mapping data indicated that it was a separate enzyme rather than a mixture of heparinase and heparitinase/Hep III. When used in conjunction with the described separation procedures, these enzymes are powerful reagents for the structural/sequence analysis of heparin and heparan sulfate.  相似文献   

11.
Many of the biological functions of heparan sulfate (HS) proteoglycans can be attributed to specialized structures within HS moieties, which are thought to modulate binding and function of various effector proteins. Cyclophilin B (CyPB), which was initially identified as a cyclosporin A-binding protein, triggers migration and integrin-mediated adhesion of peripheral blood T lymphocytes by a mechanism dependent on interaction with cell surface HS. Here we determined the structural features of HS that are responsible for the specific binding of CyPB. In addition to the involvement of 2-O,6-O, and N-sulfate groups, we also demonstrated that binding of CyPB was dependent on the presence of N-unsubstituted glucosamine residues (GlcNH2), which have been reported to be precursors for sulfation by 3-O-sulfotransferases-3 (3-OST-3). Interestingly, 3-OST-3B isoform was found to be the main 3-OST isoenzyme expressed in peripheral blood T lymphocytes and Jurkat T cells. Moreover, down-regulation of the expression of 3-OST-3 by RNA interference potently reduced CyPB binding and consequent activation of p44/42 mitogen-activated protein kinases. Altogether, our results strongly support the hypothesis that 3-O-sulfation of GlcNH2 residues could be a key modification that provides specialized HS structures for CyPB binding to responsive cells. Given that 3-O-sulfation of GlcNH2-containing HS by 3-OST-3 also provides binding sites for glycoprotein gD of herpes simplex virus type I, these findings suggest an intriguing structural linkage between the HS sequences involved in CyPB binding and viral infection.  相似文献   

12.
Heparan sulfate d-glucosaminyl 3-O-sulfotransferases (3-OSTs) catalyze the transfer of sulfate from 3'-phosphoadenosine 5'-phosphosulfate (PAPS) to position 3 of the glucosamine residue of heparan sulfate and heparin. A sixth member of the human 3-OST family, named 3-OST-5, was recently reported (Xia, G., Chen, J., Tiwari, V., Ju, W., Li, J.-P., Malmstrom, A., Shukla, D., and Liu, J. (2002) J. Biol. Chem. 277, 37912-37919). In the present study, we cloned putative catalytic domain of the human 3-OST-5 and expressed it in insect cells as a soluble enzyme. Recombinant 3-OST-5 only exhibited sulfotransferase activity toward heparan sulfate and heparin. When incubated heparan sulfate with [35S]PAPS, the highest incorporation of35S was observed, and digestion of the product with a mixture of heparin lyases yielded two major35S-labeled disaccharides, which were determined as DeltaHexA-GlcN(NS,3S,6S) and DeltaHexA(2S)-GlcN(NS,3S) by further digestion with 2-sulfatase and degradation with mercuric acetate. However, when used heparin as acceptor, we identified a highly sulfated disaccharide unit as a major product. This had a structure of DeltaHexA(2S)-GlcN(NS,3S,6S). Quantitative real-time PCR analysis revealed that 3-OST-5 was highly expressed in fetal brain, followed by adult brain and spinal cord, and at very low or undetectable levels in the other tissues. Finally, we detected a tetrasulfated disaccharide unit in bovine intestinal heparan sulfate. To our knowledge, this is the first report to describe not only the natural occurrence of tetrasulfated disaccharide unit but also the enzymatic formation of this novel structure.  相似文献   

13.
Capillary zone electrophoresis (CZE) was used to separate eight commercial disaccharide standards of the structure delta UA2X(1----4)-D-GlcNY6X (where delta UA is 4-deoxy-alpha-L-threo-hex-4-enopyranosyluronic acid, GlcN is 2-deoxy-2-aminoglucopyranose, S is sulfate, Ac is acetate, X may be S, and Y is S or Ac). These eight disaccharides had been prepared from heparin, heparan sulfate, and derivatized heparins. A similar CZE method was recently reported for the analysis of eight chondroitin and dermatan sulfate disaccharides (A. Al-Hakim and R.J. Linhardt, Anal. Biochem. 195, 68-73, 1991). Two of the standard heparin/heparan sulfate disaccharides, having an identical charge of -2, delta UA2S(1----4)-D-GlcNAc and delta UA(1----4)-D-GlcNS, were not fully resolved using standard sodium borate/boric acid buffer. This buffer had proven effective in separating chondroitin/dermatan sulfate disaccharides of identical charge. Resolution of these two heparin/heparan sulfate disaccharides could be improved by extending the capillary length, preparing the buffer in 2H2O, or eliminating boric acid. Baseline resolution was achieved in sodium dodecyl sulfate in the absence of buffer. The structure and purity of each of the eight new commercial heparin/heparan sulfate disaccharide standards were confirmed using fast-atom-bombardment mass spectrometry and high-field 1H-NMR spectroscopy. Heparin and heparan sulfate were then depolymerized using heparinase (EC 4.2.2.7), heparin lyase II (EC 4.2.2.-), heparinitase (EC 4.2.2.8), and a combination of all three enzymes. CZE analysis of the products formed provided a disaccharide composition of each glycosaminoglycan. As little as 50 fmol of disaccharide could be detected by ultraviolet absorbance.  相似文献   

14.
15.
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   

16.
In the preceding two papers, we described two new classes of sulfated N-linked oligosaccharides isolated from total cellular 35SO4-labeled macromolecules of different mammalian cell lines. The first class carries various combinations of sialic acids and 6-O-sulfate esters on typical complex-type chains, while the second carries heparin and heparan-like sequences. In this study, we have characterized a sulfophosphoglycoprotein of 140 kDa from FG-Met-2 pancreatic cancer cells whose oligosaccharides share some properties of both these classes. The molecule was localized to the cell surface by electron microscopy using a monoclonal antibody (S3-53) and by cell surface 125I-labeling. Metabolic labeling of the cells with radioactive glucosamine, methionine, inorganic sulfate, or phosphate all demonstrated a single 140-kDa molecule. Pulse-chase analysis and tunicamycin treatment indicated the glycosylation of a putative primary translation product of 110 kDa via an intermediate (120 kDa) to the mature form (140 kDa). Digestion with peptide:N-glycosidase F (PNGaseF) indicated a minimum of four N-linked glycosylation sites. PNGaseF released more than 90% of the [6-3H]GlcNH2 label and 40-70% of 35SO4 label from the immunoprecipitated 140-kDa molecule. The isolated oligosaccharides were characterized as described in the preceding two papers. The majority of [6-3H]GlcNH2-labeled molecules were susceptible to neuraminidase. More than 50% of the 35SO4 label was associated with only 5-10% of the 3H-labeled chains. Some of the sulfated chains were partly sialylated molecules with four to five negative charges. Treatment with nitrous acid released about 25% of the 35SO4 label as free sulfate, together with 6% of the [6-3H]GlcNH2 label, indicating the presence of N-sulfated glucosamine residues. Some of these oligosaccharides were degraded by heparinase and heparitinase. Therefore, while they are not as highly charged as typical heparin or heparan chains, they must share structural features that permit recognition by the enzymes. Thus, this 140-kDa glycoprotein contains at least four asparagine-linked chains substituted with a heterogeneous mixture of sulfated sequences. The heterogeneity of these molecules is as extensive as that described for whole-cell sulfated N-linked oligosaccharides in the preceding two papers.  相似文献   

17.
The antler is the most rapidly growing tissue in the animal kingdom. According to previous reports, antler glycosaminoglycans (GAGs) consist of all kinds GAGs except for heparan sulfate (HS). Chondroitin sulfate is the major antler GAG component comprising 88% of the total uronic acid content. In the current study, we have isolated HS from antler for the first time and characterized it based on both NMR spectroscopy and disaccharide composition analysis. Antler GAGs were isolated by protease treatment and followed by cetylpyridinium chloride precipitation. The sensitivity of antler GAGs to heparin lyase III showed that this sample contained heparan sulfate. After incubation of antler GAGs with chondroitin lyase ABC, the HS-containing fraction was recovered by ethanol precipitation. The composition of HS disaccharides in this fraction was determined by its complete depolymerization with a mixture of heparin lyase I, II, and III and analysis of the resulting disaccharides by the reversed-phase (RP) ion pairing-HPLC, monitored by the fluorescence detection using 2-cyanoacetamide as a post-column labeling reagent. Eight unsaturated disaccharides (DeltaUA-GlcNAc, DeltaUA-GlcNS, DeltaUA-GlcNAc6S, DeltaUA2S-GlcNAc, DeltaUA-GlcNS6S, DeltaUA2S-GlcNS, DeltaUA2S-GlcNAc6S, DeltaUA2S-GlcNS6S) were produced from antler HS by digestion with the mixture of heparin lyases. The total content of 2-O-sulfo disaccharide units in antler HS was higher than that of heparan sulfate from most other animal sources.  相似文献   

18.
Heparin like glycosaminoglycans (HLGAGs) are struc-turally complex linear polysaccharides composed of re-peating disaccharide unit of uronic (α-L-iduronic or β-D-glucuronic) acid linked 1→4 to α-D-glucosamine, whichis a highly variable sulfation pattern and ascribes to eachglycosaminoglycan (GAG) chain a unique structuralsignature. This signature dictates specific the GAG-pro-tein interactions underlying critical biological processesrelated to cell and tissue functions [1]. Only in fe…  相似文献   

19.
The disaccharide repeating-units of heparan sulfate   总被引:11,自引:0,他引:11  
Five disaccharides have been isolated after degradation of heparan sulfate by heparinase (heparin lyase) and heparitinase (heparan sulfate lyase) and are suggested to represent the repeating units of the polysaccharide. They all contain a 4,5-unsaturated uronic acid residue and are: (a) A trisulfated disaccharide that is apparently identical to a disaccharide repeating-unit of heparin; (b) a disulfated disaccharide that seems unique for heparan sulfate and contains 2-deoxy-2-sulfamidoglucose and uronic acid sulfate residues; (c) a nonsulfated disaccharide containing a 2-acetamido-2-deoxyglucose residue; (d) a monosulfated disaccharide containing a 2-acetamido-2-deoxyglucose sulfate residue; and (e) a monosulfated disaccharide containing a 2-deoxy-2-sulfamidoglucose residue. Yields of these disaccharides from different heparan sulfate fractions are discussed in relation to possible arrangements of these units in the intact polymer.  相似文献   

20.
Abstract

The structure of heparinase II/III belonging to family 12 polysaccharide lyase (PsPL12a) from Pseudopedobacter saltans was generated by homology modeling. Multiple sequence alignment showed conserved (Asn216, Tyr270 and His400) and semi-conserved active site amino acid residues. The modeled structure of PsPL12a displayed α/α toroid domain at N-terminal and antiparallel β sheets at C-terminal domain. The modeled structure was similar to those of heparinases from polysaccharide lyase 12 and 21 families. Validation of PsPL12a model by Ramachandran plot showed 94.6% of residues in the favored region, 5.2% of residues in the allowed region and only 0.2% of residues in the outlier region. The area and volume computed for PsPL12a displayed nearly a closed conformation of the active site, similar to HepIII from Bacteroides thetaiotaomicron. The charge calculation on the surface of the PsPL12a structure showed the higher distribution of positive charge in the active site cleft as compared with other homologous structures. Molecular docking study of MD-simulated PsPL12a structure with heparin oligosaccharide showed high binding affinity as compared with heparan sulfate oligosaccharides. Comparison of the active site of modeled PsPL12a with other homologous heparinases revealed putative catalytic triad involving the residues Asn216, His400 and Tyr270. Small-angle X-ray scattering analysis of PsPL12a displayed a fully folded and boxing glove-like envelop.

Communicated by Ramaswamy H. Sarma  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