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 0 毫秒
1.
The total degradation of heparin by the joint action of a purified heparinase and a heparitinase from Flavobacterium heparinum is reported. The heparinase acts directly upon heparin, yielding 52% of a trisulfated disaccharide (O-(alpha-L-ido-4-enepyranosyluronic acid 2-sulfate)-(1leads to 4)-2sulfoamino-2-deoxy-D-glucose 6-sulfate) and 40% of a tetrasaccharide besides small amounts of hexa- and disaccharides. The tetrasaccharide is in turn completely degraded by the heparitinase, forming trisulfated disaccharide and disulfated disaccharide (O-(alpha-D-glyco-4-enepyranosyluronic acid)-(1leads to 4)-2-sulfoamino-2-deoxy-D-glucose 6-sulfate) in equal amounts. These and other results indicate that the tri- and disulfated disaccharides are linked alternately, in a proportion of 3:1, respectively. The primary structure of heparin and the mode of action of the heparinase and the heparitinase are proposed based on the analysis of the different products formed by the action of the enzymes.  相似文献   

2.
In the previous paper (Myette, J. R., Shriver, Z., Claycamp, C., McLean, M. W., Venkataraman, G., and Sasisekharan, R. (2003) J. Biol. Chem. 278, 12157-12166), we described the molecular cloning, recombinant expression, and preliminary biochemical characterization of the heparin/heparan sulfate 2-O-sulfatase from Flavobacterium heparinum. In this paper, we extend our structure-function investigation of the 2-O-sulfatase. First, we have constructed a homology-based structural model of the enzyme active site, using as a framework the available crystallographic data for three highly related arylsulfatases. In this model, we have identified important structural parameters within the enzyme active site relevant to enzyme function, especially as they relate to its substrate specificity. By docking various disaccharide substrates, we identified potential structural determinants present within these substrates that would complement this unique active site architecture. These determinants included the position and number of sulfates present on the glucosamine, oligosaccharide chain length, the presence of a Delta4,5-unsaturated double bond, and the exolytic versus endolytic potential of the enzyme. The predictions made from our model provided a structural basis of substrate specificity originally interpreted from the biochemical and kinetic data. Our modeling approach was further complemented experimentally using peptide mapping in tandem with mass spectrometry and site-directed mutagenesis to physically demonstrate the presence of a covalently modified cysteine (formylglycine) within the active site. This combinatorial approach of structure modeling and biochemical studies provides insight into the molecular basis of enzyme function.  相似文献   

3.
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   

4.
5.
The capsular polysaccharide from E. Coli, strain K5 composed of ...-->4)beta-D-GlcA(1-->4)alpha-D-GlcNAc(1-->4)beta-D-GlcA (1-->..., chemically modified K5 polysaccharides, bearing sulfates at C-2 and C-6 of the hexosamine moiety and at the C-2 of the glucuronic acid residues as well as 2-O desulfated heparin were used as substrates to study the specificity of heparitinases I and II and heparinase from Flavobacterium heparinum. The natural K5 polysaccharide was susceptible only to heparitinase I forming deltaU-GlcNAc. N-deacetylated, N-sulfated K5 became susceptible to both heparitinases I and II producing deltaU-GlcNS. The K5 polysaccharides containing sulfate at the C-2 and C-6 positions of the hexosamine moiety and C-2 position of the glucuronic acid residues were susceptible only to heparitinase II producing deltaU-GlcNS,6S and deltaU,2S-GlcNS,6S respectively. These combined results led to the conclusion that the sulfate at C-6 position of the glucosamine is impeditive for the action of heparitinase I and that heparitinase II requires at least a C-2 or a C-6 sulfate in the glucosamine residues of the substrate for its activity. Iduronic acid-2-O-desulfated heparin was susceptible only to heparitinase II producing deltaU-GlcNS,6S. All the modified K5 polysaccharides as well as the desulfated heparin were not substrates for heparinase. This led to the conclusion that heparitinase II acts upon linkages containing non-sulfated iduronic acid residues and that heparinase requires C-2 sulfated iduronic acid residues for its activity.  相似文献   

6.
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   

7.
8.
A sulphamidase and a sulphoesterase were isolated from adapted cells of Flavobacterium heparinum. These enzymes were partially purified from the ;heparinases' present in the bacterial extracts and characterized. The sulphamidase has a high specificity for glucosamine N-sulphate and glucosamine 2,6-disulphate. The activity decreases sharply with increasing molecular weight of the substrates tested. The sulphamidase and the sulphoesterase activities were distinguished from each other by their different sensitivities to concentration of phosphate ion and to temperature. The importance of these enzymes in the study of the structure of heparin is discussed.  相似文献   

9.
The soil bacterium Flavobacterium heparinum produces several enzymes that degrade heparan sulfate glycosaminoglycans (HSGAGs) in a sequence-specific manner. Among others, these enzymes include the heparinases and an unusual glycuronidase that hydrolyzes the unsaturated Delta4,5 uronic acid at the nonreducing end of oligosaccharides resulting from prior heparinase eliminative cleavage. We report here the molecular cloning of the Delta4,5 glycuronidase gene from the flavobacterial genome and its recombinant expression in Escherichia coli as a highly active enzyme. We also report the biochemical and kinetic characterization of this enzyme, including an analysis of its substrate specificity. We find that the Delta4,5 glycuronidase discriminates on the basis of both the glycosidic linkage and the sulfation pattern within its saccharide substrate. In particular, we find that the glycuronidase displays a strong preference for 1-->4 linkages, making this enzyme specific to heparin/heparan sulfate rather than 1-->3 linked glycosaminoglycans such as chondroitin/dermatan sulfate or hyaluronan. Finally, we demonstrate the utility of this enzyme in the sequencing of heparinase-derived HSGAG oligosaccharides.  相似文献   

10.
11.
The analyses of the products formed from heparitin sulfates by the action of two heparitinases and a heparinase from Flavorbacterium heparinum is reported. Heparitin sulfates A and B are degraded by heparitinase I yielding two disaccharides, one of them composed of N-acetylucosamine and an unsaturated uronic, joined by alpha(1 lead to 4) linkage, and the other, with the same composition but with an O-sulfate at the hexosamine moiety. A third disaccharide is also formed from heparitin sulfate B, by the action of the same enzyme, composed of glucosamine N-sulfate and an unsaturated uronic acid joined probably by alpha(1 lead to 4) linkage. Besides these three disaccharides, heparitin sulfate B yields, by the action of heparitinase I, an oligosaccharide (with an average molecular weight of 6000) which is completely degraded by the heparitinase II yielding a disaccharide composed of glucosamine 2,6-disulfate and unsaturated uronic acid. All the disaccharides are further degraded by alpha-glycuronidase from Flavobacterium heparinum yielding the respective monosaccharides. Based on these and other analyses the possible structures of the heparitin sulfates are proposed.  相似文献   

12.
The biological actions of heparin and heparan sulfate, two structurally related glycosaminoglycans, depend on the organization of the complex heparanome. Due to the structural complexity of the heparanome, the sequence of variably sulfonated uronic acid and glucosamine residues is usually characterized by the analysis of smaller oligosaccharide and disaccharide fragments. Even characterization of smaller heparin and heparan sulfate oligosaccharide or disaccharide fragments using simple 1D 1H NMR spectroscopy is often complicated by the extensive signal overlap. 13C NMR signals, on the other hand, overlap less and therefore, 13C NMR spectroscopy can greatly facilitate the structural elucidation of the complex heparanome and provide finer insights into the structural basis for biological functions. This is the first report of the preparation of anomeric carbon-specific 13C-labeled heparin and heparan sulfate precursors from the Escherichia coli K5 strain. Uniformly 13C- and 15N-labeled precursors were also produced and characterized by 13C NMR spectroscopy. Mass spectrometric analysis of enzymatically fragmented disaccharides revealed that anomeric carbon-specific labeling efforts resulted in a minor loss/scrambling of 13C in the precursor backbone, whereas uniform labeling efforts resulted in greater than 95% 13C isotope enrichment in the precursor backbone. These labeled precursors provided high-resolution NMR signals with great sensitivity and set the stage for studying the heparanome-proteome interactions.  相似文献   

13.
14.
15.
A heparinase that degrades both heparin and heparan sulfate (HS) was purified to homogeneity from the cell-free extract of Bacillus circulans HpT298. The purified enzyme had a single band on SDS-polyacrylamide gel electrophoresis with an estimated molecular mass of 111,000. The enzyme showed optimal activity at pH 7.5 and 45 degrees C, and its activity was stimulated in the presence of 5 mM CaCl2, BaCl2, or MgCl2. Analysis of substrate specificity and degraded disaccharides demonstrated that the enzyme acts on both heparin and HS, similar to heparinase II from Flavobacterium heparinum.  相似文献   

16.
17.
Saccharides produced by the action of heparinase II on native pig mucosal heparin (heparin IS), de-N-sulphated heparin (heparin IH), N-acetylheparin (heparin IA), de-N/O-sulphated heparin (heparin IVH), de-O-sulphated heparin (heparin IVS) and de-O-sulphated N-acetylheparin (heparin IVA) were analysed by reversed-phase HPLC using Spherisorb ODS2. Fractions obtained by gel filtration with Bio-Gel P-4 were similarly examined. Heparin IS gave delta UA-2S----GlcNS-6S (IS) as the major unsaturated disaccharide and lesser amounts of delta UA----GlcNS-6S (IIS), delta UA-2S----GlcNS (IIIS), delta UA----GlcNS (IVS), delta UA-2S----GlcNAc-6S (IA), delta UA----GlcNAc-6S (IIA), delta UA-2S----GlcNAc (IIIA) and delta UA----GlcNAc (IVA). Heparins IA, IVA and IVS gave as the predominant unsaturated disaccharide that corresponding to the major repeat structure of the polymer. These were respectively delta UA-2S----GlcNAc-6S (IA), delta UA-GlcNAc (IVA) and delta UA----GlcNS (IVS). Minor disaccharides from the heterogeneous structure in native pig heparin and from residual O-sulphates after the de-O-sulphating process were detected. Heparin IH was degraded more slowly than any of the N-substituted heparins. The predominant unsaturated disaccharide was IH, which was derived from the major repeating unit. In addition, disaccharides IIH, IIIH, IA, IIA and IVA were detected. Heparin IVH showed little degradation, the unsaturated disaccharide IVH not being detected after 24 h. Disaccharide IVA was obtained from the heterogeneous sequence in heparin IVH. Several higher oligosaccharides were identified in the gel-filtration fractions including saccharides from the linkage region (for heparin IS and IVA) and the anti-thrombin binding site (for heparin IS only). A tetrasaccharide and hexasaccharide, with the structures delta UA----GlcNAc----UA----GlcNAc and delta UA----GlcNAc----UA----GlcNAc----UA----GlcNAc, were present in the HPLC profiles of heparins IA and IVA.  相似文献   

18.
Heparitinase [EC 4.2.2.8, heparitin sulfate lyase] was prepared from an extract of cultured cells of Flavobacterium heparinum. Purification of the enzyme was achieved by repeating the hydroxyapatite column chromatography. The enzyme was used to degrade heparan sulfate occurring on the surfaces of ascites hepatoma cells, AH 66. From the supernatant of the enzyme-treated cells, breakdown products from heparan sulfate could be detected by paper chromatography. The heparitinase was found to be more effective than trypsin in removing heparan sulfate from the cells. Furthermore, on analyzing glycosaminoglycans and glycopeptides from the enzyme-treated cells and control cells, it was concluded that heparan sulfate was exclusively present on the cell surface and accessible to the heparitinase whereas other cell surface complex carbohydrates remained intact.  相似文献   

19.
Kurimoto E  Nishi Y  Yamaguchi Y  Zako T  Iizuka R  Ide N  Yohda M  Kato K 《Proteins》2008,70(4):1257-1263
Group II chaperonin (CPN) cooperates with prefoldin (PFD), which forms a jellyfish-shaped heterohexameric complex with a molecular mass of 87 kDa. PFD captures an unfolded protein with the tentacles and transfers it to the cavity of CPN. Although X-ray crystal structures of CPN and PFD have been reported, no structural information has been so far available for the terminal regions of the PFD tentacles nor for the C-terminal segments of CPNs, which were regarded to be functionally significant in the previous studies. Here we report 13C NMR analyses on archaeal PFD, CPN, and their complex, focusing on those structurally uncharacterized regions. The PFD and CPN complexes selectively labeled with 13C at methionyl carbonyl carbons were separately and jointly subjected to NMR measurements. 13C NMR spectral data demonstrated that the N-terminal segment of the alpha and beta subunits of PFD as well as the C-terminal segments of the CPN hexadecamer retain significant degrees of freedom in internal motion even in the complex with a molecular mass of 1.1 MDa.  相似文献   

20.
T J Williams  R W Henkens 《Biochemistry》1985,24(10):2459-2462
Using 13C NMR spectroscopy, we have further investigated the binding of HCO3- in the active site of an artificial form of human carbonic anhydrase I in which the native zinc is replaced by Co(II). The Co(II) enzyme, unlike all other metal-substituted derivatives, has functional properties closely similar to those of the native zinc enzyme. By measuring the spin-lattice relaxation rate and the line width for both the CO2 and HCO3- at two field strengths, we have determined both the paramagnetic effects that reflect substrate binding and the exchange effects due to catalysis at chemical equilibrium. The following are the results at 14 degrees C and pH 6.3 (1) HCO3- is bound in the active site of the catalytically competent enzyme with the 13C of the HCO3- located 3.22 +/- 0.02 A from the Co(II); (2) the apparent equilibrium dissociation constant for the bound HCO3- is 7.6 +/- 1.5 mM, determined by using the paramagnetic effects on the line widths, and 10 +/- 2 mM, determined by using the exchange effects; (3) the lifetime of HCO3- bound to the metal is (4.4 +/- 0.4) X 10(-5) s; (4) the overall catalyzed CO2 in equilibrium HCO3- exchange rate constant of the Co(II) enzyme is (9.6 +/- 0.4) X 10(3) s-1; (5) the electron spin relaxation time of the Co(II), determined by using paramagnetic effects on the bound HCO3-, is (1.1 +/- 0.1) X 10(-11) s. The data did not provide any direct information on the binding of CO2.(ABSTRACT TRUNCATED AT 250 WORDS)  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