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 453 毫秒
1.
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   

2.
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   

3.
The rapid preparation of multimilligram quantities of five heparin-derived oligosaccharides (1–5) is described. These oligosaccharides are the final products obtained from the action of heparin lyase (heparinase, E.C. 4.2.2.7) at its primary sites in the heparin polymer. Five oligosaccharides comprise from 75–85 wt% of commercial porcine mucosal heparins and are recovered in good yield and high purity. Four of these five oligosaccharides were further acted upon at much lower rates by prolonged treatment with heparin lyase or heparan monosulfate lyase (heparitinase, E.C. 4.2.2.8), revealing the subspecificities of these enzymes. These oligosaccharides were used as defined substrates for heparin lyase and heparan monosulfate lyase and their kinetic constants were obtained. Potential applications for these oligosaccharides include their use as defined substrates for purification of heparin monosulfate lyases, and for establishing the catalytic purity of enzyme preparations.  相似文献   

4.
Heparin lyase I has been purified from Flavobacterium heparinum and has been partially characterized (Yang, V. C., Linhardt, R. J., Berstein, H., Cooney, C. L., and Langer, R. (1985) J. Biol. Chem. 260, 1849-1857). There has been no report of the purification of the other polysaccharide lyases from this organism. Although all three of these heparin/heparan sulfate lyases are widely used, with the exception of heparin lyase I, there is no information on their purity or their physical and kinetic characteristics. The absence of pure heparin lyases and a lack of understanding of the optimal catalytic conditions and substrate specificity has stood in the way of the use of these enzymes as reagents for the specific depolymerization of heparin and heparan sulfate into oligosaccharides for structure and activity studies. This paper describes a single, reproducible scheme to simultaneously purify all three of the heparin lyases from F. heparinum to apparent homogeneity. Heparin lyase I (heparinase, EC 4.2.2.7), heparin lyase II (no EC number), and heparin lyase III (heparitinase, EC 4.2.2.8) have molecular weights (by sodium dodecyl sulfate-polyacrylamide gel electrophoresis) and isoelectric points (by isoelectric focusing) of M(r) 42,800, pI 9.1-9.2, M(r) 84,100, pI 8.9-9.1, M(r) 70,800, pI 9.9-10.1, respectively. Their amino acid analyses and peptide maps demonstrate that while these proteins are different gene products they are closely related. The kinetic properties of the heparin lyases have been determined as well as the conditions to optimize their activity and stability. These data should improve the application of these important enzymes in the study of heparin and heparan sulfate.  相似文献   

5.
Heparinase and heparitinase were separated from an extract of Flavobacterium heparinum, induced with heparin by using column chromatography on hydroxylapatite. As the heparinase preparation contained chondroitinases B and C, chondroitinase B was removed by rechromatography on a hydroxylapatite column. Chondroitinase C was then eliminated by column chromatography on O-phosphono(“phospho”)-cellulose. The heparinase preparation thus obtained was free from sulfoamidase for 2-deoxy-2-sulfoamino-D-glucose (GlcN-2S), sulfatase for 2-amino-2-deoxy-6-O-sulfo D-glucose (GlcN-6S), as well as (δ4,5glycosiduronase for the unsaturated disaccharides obtained from heparin. The remaining sulfatase for 4-deoxy-α-L-thero-hex-4-enopyranosyluronic acid 2-sulfate (δUA-2S) in the heparinase preparation was removed by affinity chromatography with dermatan sulfate-bound AH-Sepharose 4B coated with dermatan sulfate. The heparitinase preparation separated by column chromatography on hydroxyla patite was purified by affinity chromatography with heparin-bound AH-Sepharose 4B coated with heparin. Sulfatase for 2-amino-2-deoxy-6-O-sulfo-D-glucose (GlcN-6S) and δ4,5glycosiduronase for the unsaturated disaccharides obtained from heparin were removed by this chromatography. Sulfatase for 4-deoxy-α-L-threo-hex-4-enopyranosyluronic acid 2-sulfate (δUA-2S) remaining in the heparitinase preparation was finally removed by column chromatography on hydroxylapatite. The recoveries of the purified preparations of heparinase and heparitinase were estimated to be 39 and 50%, respectively, from the crude enzyme fractions obtained by the first column chromatography on hydroxyl- patite. The purified heparinase and heparitinase were free from all enzymes that could degrade the sulfated unsaturated disaccharides produced from heparin with heparinase.  相似文献   

6.
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   

7.
Porcine intestinal mucosal heparan sulfate was exhaustivelydepolymerized on a large scale using beparin lyase II (heparinaseII) or heparin lyase III (heparitinase, EC 4.2.2.8 [EC] ). The oligosaccharidemixtures formed with each enzyme were fractionated by low pressuregel permeation chromatography. Size-uniform mixtures of disaccharides,tetrasaccharides, and hexasaccharides were obtained. Each size-fractionatedmixture was then purified on the basis of charge by repetitivesemipreparative strong-anion-exchange high-performance liquidchromatography. This approach has led to the isolation of 13homogenous oligosaccharides. The purity of each oligosaccharidewas demonstrated by the presence of a single peak on analyticalstrong-anion-exchange high-performance liquid chromatographyand reversed polarity capillary electrophoresis. The structuresof these oligosaccharides were established using 500 MHz one-and two-dimensional nuclear magnetic resonance spectroscopy.Three of the thirteen structures that were solved were novelwhile the remaining 10 have been previously described. All ofthe structures obtained using heparin lyase III contained a  相似文献   

8.
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   

9.
  • 1.1. The disaccharide sequences of a heparan sulfate isolated from Anomantidae sp. was determined with the aid of heparitinase I, heparitinase II from Flavobacterium heparinum, mollusc β-glucuronidase and α-N-acetylglucosaminidase besides nitrous acid degradation and chemical analyses.
  • 2.2. Like the mammalian heparan sulfates the mollusc heparan sulfate is composed of different oligosaccharide blocks of N-acetylated disaccharides, N-sulfated disaccharides and N,6-sulfated disaccharides and has in its nonreducing end the monosaccharide glucosamine 2,6-disulfate.
  • 3.3. The oligosaccharides produced by heparitinase I degradation contain at their reducing ends a N-acetylated, 6-sulfated disaccharide.
  • 4.4. These and other results lead to the conclusion that the general structure of the heparan sulfate is maintained through evolution.
  相似文献   

10.
The structural properties of fibroblast heparan sulfate (HS) that are necessary for it to bind strongly to basic fibroblast growth factor (bFGF) have been investigated using bFGF affinity chromatography. Specific enzymic and chemical scission of HS, together with chemical N-desulfation, revealed that N-sulfate groups and iduronate-2-sulfates (IdoA(2-OSO3)) were essential for the interaction. bFGF-affinity chromatography of sulfated oligosaccharides released from HS by treatment with heparitinase led to the identification of an oligosaccharide component (oligo-H), seven disaccharides in length, with a similar affinity for bFGF as the parent molecule. Heparinase treatment of this fraction abolished the high affinity binding to bFGF. Analysis of oligo-H indicated that 74% of the disaccharide units had the structure IdoA(2-OSO3)alpha 1,4GlcNSO3; the remainder comprised N-acetylated and N-sulfated units, the majority of which were devoid of O-sulfate groups. Oligo-H was fully degraded to disaccharides by treatment with nitrous acid. These results indicate that the sequence of oligo-H is as shown below. delta GlcA beta 1,4GlcNSO3 alpha 1,4[IdoA(2-OSO3)alpha 1,4GlcNSO3]5 alpha 1, 4IdoA alpha 1,4GlcNAc Sulfated oligosaccharides of similar size but with a lower affinity for bFGF had a reduced concentration of IdoA(2-OSO3) but significant quantities of GlcNSO3(6-OSO3) and GlcNAc(6-OSO3). The data indicate a primary role for contiguous sequences of IdoA(2-OSO3)alpha 1,4GlcNSO3 in mediating the high affinity binding between fibroblast HS and bFGF.  相似文献   

11.
Chen M  Bridges A  Liu J 《Biochemistry》2006,45(40):12358-12365
Heparan sulfate plays a wide range of physiological and pathological roles. Heparan sulfate consists of glucosamine and glucuronic/iduronic acid repeating disaccharides with various sulfations. Synthesis of structurally defined heparan sulfate oligosaccharides remains a challenge. Access to nonsulfated and unepimerized heparan sulfate backbone structures represents an essential step toward de novo enzymatic synthesis of heparan sulfate. The nonsulfated, unepimerized backbone heparan sulfate is similar to the capsular polysaccharide from Escherichia coli strain K5. The biosynthesis of this capsular polysaccharide involves in N-acetylglucosaminyltransferase (KfiA) and d-glucuronyltransferase (KfiC). In this study, we report the characterization of purified KfiA. KfiA was expressed in a C-terminal six-His fusion protein in BL21 star cells coexpressing chaperone proteins GroEL and GroES. The recombinant KfiA was purified to homogeneity with a Ni-agarose column. The binding affinities of various UDP-sugars for KfiA were determined using isothermal calorimetry titration, indicating that both the N-acetyl group and sugar type may be essential for donor substrates to bind KfiA. Kinetic analysis of KfiA toward different sizes of oligosaccharide revealed that KfiA is less sensitive to the size of the acceptor substrates. The results from this study open a new approach for the synthesis of the heparan sulfate backbone.  相似文献   

12.
A new method that we have called 'oligosaccharide mapping' is described for the analysis of radiolabelled heparan sulphate and other glycosaminoglycans. The method involves specific enzymic or chemical scission of polysaccharide chains followed by high-resolution separation of the degradation products by polyacrylamide-gradient-gel electrophoresis. The separated oligosaccharides are immobilized on charged nylon membranes by electrotransfer and detected by fluorography. A complex pattern of discrete bands is observed covering an oligosaccharide size range from degree of polymerization (d.p.) 2 (disaccharide) to approximately d.p. 40. Separation is due principally to differences in Mr, though the method also seems to detect variations in conformation of oligosaccharide isomers. Resolution of oligosaccharides is superior to that obtained with isocratic polyacrylamide-gel-electrophoresis systems or gel chromatography, and reveals structural details that are not accessible by other methods. For example, in this paper we demonstrate a distinctive repeating doublet pattern of iduronate-rich oligosaccharides in heparitinase digests of mouse fibroblast heparan sulphate. This pattern may be a general feature of mammalian heparan sulphates. Oligosaccharide mapping should be a valuable method for the analysis of fine structure and sequence of heparan sulphate and other complex polysaccharides, and for making rapid assessments of the molecular distinctions between heparan sulphates from different sources.  相似文献   

13.
The carbohydrate antigen on heparan sulfate recognized by monoclonal antibody 10E4 is uniquely codistributed with the abnormal prion protein, PrP(Sc), even in the earliest detectable brain lesions of scrapie-infected mice. Determining the chemical structure of 10E4 antigen is, therefore, an important aspect of structure elucidation of scrapie lesions, and a prerequisite for designing experiments to understand its role in scrapie pathogenesis. Toward this aim, we have examined preparations of heparan sulfate, with differing sulfate contents, for binding by 10E4 antibody. The highest antigenicity was observed in a preparation (HS-1) with the lowest sulfate content. HS-1 was partially depolymerized with heparin lyase III, and oligosaccharide fragments examined for 10E4 antigen expression by the neoglycolipid technology. An antigen-positive and two antigen-negative tetrasaccharides were isolated and examined by electrospray mass spectrometry. The antigen-positive tetrasaccharide sequence on heparan sulfate was thus deduced to contain a unique unsulfated motif that includes an N-unsubstituted glucosamine in the sequence, UA-GlcN-UA-GlcNAc. Antibody binding experiments with neoglycolipids prepared from a series of heparin/heparan sulfate disaccharides, and the trisaccharide derived from the antigen-positive tetrasaccharide after removal of the terminal hexuronic acid, show that both the penultimate glucosamine and the outer nonsulfated hexuronic acid are important for 10E4 antigenicity.  相似文献   

14.
In order to prepare a series of N-acetylheparosan (NAH)-related oligosaccharides, bacterial NAH produced in Escherichia coli strain K5 was partially depolymerized with heparitinase I into a mixture of even-numbered NAH oligosaccharides, having an unsaturated uronic acid (DeltaUA) at the non-reducing end. A mixture of odd-numbered oligosaccharides was derived by removing this DeltaUA in the aforementioned mixture by a 'trimming' reaction using mercury(II) acetate. Each oligosaccharide mixture was subjected to gel-filtration chromatography to generate a series of size-uniform NAH oligosaccharides of satisfactory purity (assessed by analytical anion-exchange HPLC), and their structures were identified by MALDITOF-MS, ESIMS, and 1H NMR analysis. As a result, a microscale preparation of a series of both even- and odd-numbered NAH oligosaccharides was achieved for the first time. The developed procedure is simple and systematic, and thus, should be valuable for providing not only research tools for heparin/heparan sulfate-specific enzymes and their binding proteins, but also precursor substrates with medical applications.  相似文献   

15.
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   

16.
Topically applied heparin and heparan sulfate disaccharides, with the basic structure delta-4,5 uronyl-(1----4)-glucosamine and bearing a sulfate at the C-6 position of the glucosamine residue, are antihemostatics as potent as heparin, producing uncontrollable hemorrhage from small blood vessels. The finding that other sulfated disaccharides with the same sulfate:hexosamine:uronic acid ratios but with the sulfate at a different position (C-2), or with different glycosidic linkage (1----3), were inactive as inhibitors of hemostasis indicates that a specific structure is needed to produce the effect. The inhibitory activity of the normal hemostatic process could be reversed by ATP. Molecular models show that part of the disaccharide inhibitors and ATP hold a similar structural conformation.  相似文献   

17.
Pharmaceutical heparin and heparan sulfate, isolated from a side-stream of a commercial heparin manufacturing process, have been enzymatically depolymerzed with heparin lyases obtained from Flavobacterium heparinun. Heparin afforded a trisulfated disaccharide product that was recovered from the reaction mixture using gel permeation chromatography. Heparan sulfate afforded unsulfated disaccharide that was conveniently recovered from the product mixture by ion exchange chromatography. Both disaccharides were obtained in gram amounts at 90% or higher purity. Both enzymatically prepared disaccharides were chemically protected to prepare building blocks required for the future chemical synthesis of therapeutically valuable heparin oligosaccharides.  相似文献   

18.
Oligosaccharides prepared from glycosaminoglycans (GAGs) including heparin, heparan sulfate, chondroitin sulfates, dermatan sulfate, and keratan sulfate were analyzed using reverse-phase ion-pairing HPLC and ion-exchange HPLC with suppressed conductivity detection. The results were compared with those obtained by strong anion-exchange HPLC using uv detection. These oligosaccharides were first prepared by enzymatically depolymerizing the GAGs with enzymes including heparin lyase (EC 4.2.2.7), heparan sulfate lyase (EC 4.2.2.8), chondroitin ABC lyase (EC 4.2.2.4), and keratan sulfate hydrolase (EC 3.2.1.103). Analysis was then performed without derivitization under isocratic conditions with a limit of sensitivity in the picomole range. Preliminary studies suggest that this approach may be particularly useful in examining oligosaccharides having no uv chromophore such as those prepared from keratan sulfate.  相似文献   

19.
Heparitinase I, a key lyase enzyme essential for structural analysis of heparan sulfate (HS), degrades HS domains that are undersulfated at glucuronyl residues through an elimination mechanism. Earlier studies employed viscosimetric measurements and electrophoresis to deduce the mechanism of action of heparitinase I and two other related lyases, heparitinase II and heparitinase III. However, these findings lack molecular evidence for the intermediates formed and could not distinguish whether the cleavage occurred from the reducing end or the nonreducing end. In the current study, 2-aminoacridone (2-AMAC)-labeled HS precursor oligosaccharides of various sizes were prepared to investigate the mechanism of heparitinase I-mediated depolymerization using sensitive and quantitative methodologies. Furthermore, fluorescent (2-AMAC) tagging of HS precursor oligosaccharides allowed us to distinguish fragments that result from cleavage of the substrates at various time intervals and sites farther away from the reducing and nonreducing ends of oligosaccharide substrates. This study provides the first direct molecular evidence for a predominantly random endolytic mechanism of cleavage of HS precursor oligosaccharides by heparitinase I. This robust strategy can be adapted to deduce the mechanism of action of other heparitinases and also to deduce structural information of complex HS oligosaccharides of biological importance.  相似文献   

20.
Fragmentation of the heparan sulfate chains from bovine glomerular basement membrane (GBM) by hydrazine/nitrous acid treatment followed by NaB3H4-reduction yielded a mixture of six sulfated disaccharides containing D-glucuronic (GlcUA) or L-iduronic acid (IdUA) and terminating in 2,5-anhydro[3H]mannitol (AnManH2), in addition to the nonsulfated component GlcUA beta 1----4AnManH2. Among these products two novel disaccharide units were identified as IdUA alpha 1----4AnManH2(3-SO4) and IdUA(2-SO4)alpha 1----4AnManH2(3-SO4); these accounted for 22% of the total sulfated species indicating that there are 2-3 residues of 3-O-sulfated glucosamine/heparan sulfate chain. The disulfated disaccharide was shown through its release by direct nitrous acid treatment to be situated in a GlcNSO3-IdUA(2-SO4)-GlcNSO3(3-SO4) sequence which is distinct from that in which 3-O-sulfated glucosamine is located in the antithrombin-binding region of heparins. Analyses of heparan sulfate from lens capsule, a nonvascular basement membrane, indicated the absence of sequences containing 3-O-sulfated glucosamine, although otherwise the sulfated disaccharides produced by hydrazine/nitrous acid/Na-B3H4 treatment (GlcUA beta 1----4AnManH2(6-SO4), IdUA alpha 1----4AnManH2(6-SO4), IdUA(2-SO4)alpha 1----4AnManH2 and IdUA(2-SO4)alpha 1----4AnManH2(6-SO4] were the same as from GBM. Examination of the GBM heparan sulfate domains after nitrous acid treatment indicated that the O- as well as N-sulfate groups are clustered in an iduronic acid-rich 10-disaccharide peripheral segment, while the internal region (approximately 20 disaccharides) is composed primarily of repeating GlcUA beta 1----4GlcNAc units. The localization of chain diversity to the outer region may facilitate interactions of the heparan sulfate with other macromolecular components.  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