首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Kinetic parameters (Km and kcat.) of the two major forms (A and B) and a minor form (C) of human liver N-acetylglucosamine-6-sulphate sulphatase [Freeman, Clements & Hopwood (1987) Biochem. J. 246, 347-354] were determined with a variety of substrates matching structural aspects of the physiological substrates in vivo, namely heparin, heparan sulphate and keratan sulphate. Enzyme activity is highly specific towards glucosamine 6-sulphate or glucose 6-sulphate residues. More structurally complex substrates, in which several aspects of the aglycone structure of the natural substrate were maintained, are hydrolysed with catalytic efficiencies up to 3900 times above that observed for the monosaccharide substrate N-acetylglucosamine 6-sulphate. Forms A and B both desulphate substrates derived from keratan sulphate and heparin. Aglycone structures that influence substrate binding and/or enzyme activity were penultimate-residue 6-carboxy and 2-sulphate ester groups for heparin-derived substrates and penultimate-residue 6-sulphate ester groups for keratan sulphate-derived substrates. The 4-hydroxy group of the N-acetylglucosamine 6-sulphate or the 2-sulphaminoglucosamine 6-sulphate under enzymic attack is involved in the catalytic mechanism. The presence of a 2-amino group in place of a 2-acetamido or a 2-sulphoamino group considerably decreases the catalytic efficiency of the sulphatase, particularly in the absence of a penultimate-aglycone-residue 6-carboxy group. Both forms A and B are exo-enzymes, since activity towards internal sulphate ester bonds was not observed. The effect of incubation pH on enzyme activity towards the variety of substrates evaluated was complex and dependent on substrate aglycone structure. The presence of aglycone 2-sulphate ester, 6-carboxy group and 6-sulphate ester groups on the glucosamine 6-sulphate residue under attack considerably affects the pH response. Sulphate and phosphate ions are potent inhibitors of enzyme activity.  相似文献   

2.
Human sulphamate sulphohydrolase was purified at least 20,000-fold to homogeneity from liver with a three-step four-column procedure, which consisted of a concanavalin A-Sepharose/Blue A agarose coupled step, and Bio-Gel HT step and then a CM-Sepharose step. The procedure was also used to purify enzyme from kidney and placenta. The subunit Mr of liver, kidney and placenta sulphamate sulphohydrolase was assessed to be 56,000 by using SDS/polacrylamide-gel electrophoresis. The native protein Mr of enzyme from all three tissue sources was assessed by gel-permeation chromatography to be approx. 120,000 on Sephacryl S-300 and 100,000 on Fractogel TSK. It is probable that the native enzyme results from dimerization of subunits. Kinetic parameters (km and kcat.) of human liver sulphamate sulphohydrolase were determined with a variety of substrates matching structural aspects of the physiological substrates in vivo, namely heparin and heparan sulphate. More structurally complex substrates, in which several aspects of the aglycone structure of the natural substrate were maintained, are turned over up to 372000 times faster than the monosaccharide substrate 2-sulphaminoglucosamine. Aglycone structures that influence substrate binding and/or enzyme activity were penultimate-residue C-6 carboxy and C-2 sulphate ester groups and a post-penultimate 2-sulphaminoglucosamine residue. The C-4 hydroxy group of the 2-sulphaminoglucosamine under enzymic attack is involved in binding of substrate to enzyme. The presence of C-6 sulphate ester on the non-reducing end 2-sulphaminoglucosamine stimulates sulphamate bond hydrolysis and substrate affinity if the adjacent monosaccharide residue is idose or 2-sulphoidose, but strongly inhibits hydrolysis if the adjacent monosaccharide residue is iduronic acid. Sulphamate sulphohydrolase is an exoenzyme, since activity toward internal sulphamate bonds was not detected. The effect of incubation pH on enzyme activity towards the variety of substrates evaluated was complex and dependent on substrate aglycone structure. The presence of aglycone C-2 sulphate ester and aglycone C-6 carboxy groups and C-6 sulphate ester groups on the 2-sulphaminoglucosamine residue under attack considerably affect the pH response. Structurally complex substrates had two pH optima. Incubation temperature and buffer ionic strength markedly influenced pH optima and enzyme activity. Cu2+ and SO4(2-)ions are potent inhibitors of enzyme activity.  相似文献   

3.
The kinetic parameters (Km and V) of human arylsulphatase B (4-sulpho-N-acetylgalactosamine sulphatase) activity in cultured skin fibroblasts were determined with a variety of substrates matching structural aspects of the physiological substrates in vivo chondroitin 4-sulphate and dermatan sulphate. More structurally complex substrates, in which several aspects of the aglycone structure of the natural substrate were maintained, were desulphated up to 4400 times faster than the minimum arylsulphatase-B-specific substrate, namely the monosaccharide N-acetylgalactosamine 4-sulphate. Aglycone structures that influence substrate binding and/or enzyme activity were an adjacent-residue C-6 carboxy group and a second but internal N-acetylgalactosamine 4-sulphate residue. Arylsulphatase B activity in fibroblast homogenates assayed with O-(beta-N-acetylgalactosamine 4-sulphate)-(1----4)-O-D-(beta-glucuronic acid)-(1----3)-O-D-N-acetyl[1-3H] galactosaminitol 4-sulphate derived from chondroitin 4-sulphate as substrate clearly distinguished Maroteaux-Lamy-syndrome patients from normal controls and other mucopolysaccharidosis patients. We recommend the use of the above trisaccharide substrate for both postnatal and prenatal diagnosis of Maroteaux-Lamy syndrome.  相似文献   

4.
Glycosaminoglycans synthesized in polymorphonuclear (PMN) leucocytes isolated from blood (peripheral PMN leucocytes) and in those induced intraperitoneally by the injection of caseinate (peritoneal PMN leucocytes) were compared. Both peripheral and peritoneal PMN leucocytes were incubated in medium containing [35S]sulphate and [3H]glucosamine. Each sample obtained after incubation was separated into cell, cell-surface and medium fractions by trypsin digestion and centrifugation. The glycosaminoglycans secreted from peripheral and peritoneal PMN leucocytes were decreased in size by alkali treatment, indicating that they existed in the form of proteoglycans. Descending paper chromatography of the unsaturated disaccharides obtained by the digestion of glycosaminoglycans with chondroitinase AC and chondroitinase ABC identified the labelled glycosaminoglycans of both the cell and the medium fractions in peripheral PMN leucocytes as 55-58% chondroitin 4-sulphate, 16-19% chondroitin 6-sulphate, 16-19% dermatan sulphate and 6-8% heparan sulphate. Oversulphated chondroitin sulphate and oversulphated dermatan sulphate were found only in the medium fraction. In peritoneal PMN leucocytes there is a difference in the composition of glycosaminoglycans between the cell and the medium fractions; the cell fraction was composed of 60% chondroitin 4-sulphate, 5.5% chondroitin 6-sulphate, 16.8% dermatan sulphate and 13.9% heparan sulphate, whereas the medium fraction consisted of 24.5% chondroitin 4-sulphate, 28.2% chondroitin 6-sulphate, 33.7% dermatan sulphate and 10% heparan sulphate. Oversulphated chondroitin sulphate and oversulphated dermatan sulphate were found in the cell, cell-surface and medium fractions. On the basis of enzymic assays with chondro-4-sulphatase and chondro-6-sulphatase, the positions of sulphation in the disulphated disaccharides were identified as 4- and 6-positions of N-acetylgalactosamine. Most of the 35S-labelled glycosaminoglycans synthesized in peripheral PMN leucocytes were retained within cells, whereas those in peritoneal PMN leucocytes were secreted into the culture medium. Moreover, the amount of glycosaminoglycans in peritoneal PMN leucocytes was significantly less than that in peripheral PMN leucocytes. Assay of lysosomal enzymes showed that these activities in peritoneal PMN leucocytes were 2-fold higher than those in peripheral PMN leucocytes.  相似文献   

5.
Purified m beta-acrosin catalysed amidolysis of several p-nitroanilides with C-terminal arginine residues. Antithrombin III inhibited amidolysis catalysed by the enzyme. This effect of antithrombin III was potentiated by heparin, and to a modest extent by heparan sulphate, cellulose sulphate, dextran sulphate and xylan sulphate. De-N-sulphated heparin, de-N-sulphated N-acetylated heparin, heparin of low relative molecular mass, chondroitin 4-sulphate, chondroitin 6-sulphate, dermatan sulphate and hyaluronic acid were ineffective.  相似文献   

6.
Human glucuronate 2-sulphatase (GAS), which is involved in the degradation of the glycosaminoglycans heparan sulphate and chondroitin 6-sulphate, was purified almost 2,000,000-fold to homogeneity in 8% yield from liver with a four-step six-column procedure, which consists of a concanavalin A-Sepharose/Blue A-agarose coupled step, a DEAE-Sephacel/octyl-Sepharose coupled step, CM-Sepharose chromatography and gel-permeation chromatography. Although more than 90% of GAS activity had a pI of greater than 7.5, other forms with pI values of 5.8, 5.3, 4.7 and less than 4.0 were also present. The pI greater than 7.5 form of GAS had a native molecular mass of 63 kDa. SDS/polyacrylamide-gel-electrophoretic analysis resulted in two polypeptide subunits of molecular mass 47 and 19.5 kDa. GAS was active towards disaccharide substrates derived from heparin [O-(beta-glucuronic acid 2-sulphate)-(1----4)-O-(2,5)-anhydro[1-3H]mannitol 6-sulphate (GSMS)] and chondroitin 6-sulphate [O-(beta-glucuronic acid 2-sulphate-(1----3)-O-(2,5)-anhydro[1-3H]talitol 6-sulphate (GSTS)]. GAS activity towards GSMS and GSTS was at pH optima of 3.2 and 3.0 respectively with apparent Km values of 0.3 and 0.6 microM respectively and corresponding Vmax values of 12.8 and 13.7 mumol/min per mg of protein respectively. Sulphate and phosphate ions are potent inhibitors of enzyme activity. Cu2+ ions stimulated, whereas EDTA inhibited enzyme activity. It was concluded that GAS is required together with a series of other exoenzyme activities in the lysosomal degradation of glycosaminoglycans containing glucuronic acid 2-sulphate residues.  相似文献   

7.
Human N-acetylglucosamine-6-sulphate sulphatase was purified at least 50,000-fold to homogeneity in 78% yield from liver with a simple three-step four-column procedure, which consists of a concanavalin A-Sepharose/Blue A-agarose coupled step, chromatofocusing and Cu2+-chelating Sepharose chromatography. In all, four forms were isolated and partially characterized. Forms A and B, both with a pI greater than 9.5 and representing 30% and 60% respectively of the recovered enzyme activity, were separated by hydroxyapatite chromatography of the enzyme preparation obtained from the Cu2+-chelating Sepharose step. Both forms A and B had native molecular masses of 75 kDa. When analysed by SDS/polyacrylamide-gel electrophoresis, form A consists of a single polypeptide of molecular mass 78 kDa, whereas form B contained 48 kDa and 32 kDa polypeptide subunits. Neither form A nor form B was taken up from the culture medium into cultured human skin fibroblasts. The two other forms (C and D), with pI values of 5.8 and 5.4 respectively, represented approx. 7% and 3% of the total recovered enzyme activity. The native molecular masses of forms C and D were 94 kDa and approx. 75 kDa respectively. Form C contained three polypeptides with molecular masses of 48, 45 and 32 kDa. N-Acetylglucosamine-6-sulphate sulphatase activity was measured with a radiolabelled disaccharide substrate derived from heparin. The development of this substrate enabled the isolation and characterization of N-acetylglucosamine-6-sulphate sulphatase to proceed efficiently. Forms A, B and C had pH optima of 5.0, Km values of 11.7, 14.2 and 11.1 microM respectively and Vmax. values of 105, 60 and 53 nmol/min per mg of protein respectively. The molecular basis of the multiple forms of this sulphatase is not known. It is postulated that the differences in structure and properties of the four enzyme forms are due to differences in the state of processing of a large subunit.  相似文献   

8.
Jaya  P.  Kurup  P. A. 《Journal of biosciences》1986,10(4):487-493
Magnesium deficiency in rats has significant effect on the concentration of different glycosaminoglycans in the tissues, the nature of the change being different in different tissues. Total glycosaminoglycans, chondroitin-4-sulphate + chondroitin-6-sulphate and dermatan sulphate increased in the aorta while hyaluronic acid, heparan sulphate and heparin decreased. In the liver, total glycosaminoglycans, hyaluronic acid, chondroitin-4-sulphate + 6-sulphate and heparin decreased while total glycosamino-glycans and all the glycosaminoglycan fractions increased in the heart. In the kidney, total glycosaminoglycans showed no significant alteration, hyaluronic acid and heparin decreased while chondroitin-4-sulphate + 6-sulphate increased. Activity of biosynthetic enzymesviz. glucosamine-o-phosphate isomerase and UDPG-dehydrogenase showed decrease in the liver. The concentration of 3’-phosphoadenosine 5’-phosphosulphate, activity of sulphate activating system and sulphotransferase were also similarly altered in the liver in magnesium deficiency.  相似文献   

9.
Action pattern of polysaccharide lyases on glycosaminoglycans   总被引:2,自引:1,他引:1  
The action pattern of polysaccharide lyases on glycosaminoglycansubstrates was examined using viscosimetric measurements andgradient polyacrylamide gel electrophoresis (PAGE). Heparinlyase I (heparinase, EC 4.2.2.7 [EC] ) and heparin lyase II (no ECnumber) both acted on heparin in a random endolytic fashion.Heparin lyase II showed an ideal endolytic action pattern onheparan sulphate, while heparin lyase I decreased the molecularweight of heparan sulphate more slowly. Heparin lyase III (heparitinase,EC 4.2.2.8 [EC] ) acted endolytically only on heparan sulphate anddid not cleave heparin. Chondroitin ABC lyase (chondroitinaseABC, EC 4.2.2.4 [EC] ) from Proteus vulgaris acted endolytically onchondroitin-6-sulphate (chondroitin sulphate C) and dermatansulphate at nearly identical initial rates, but acted on chondroitin-4-sulphate(chondroitin sulphate A) at a reduced rate, decreasing its molecularweight much more slowly. Two chondroitin AC lyases (chondroitinaseAC, both EC 4.2.2.5 [EC] ) were examined towards chondroitin-4- and-6-sulphates. The exolytic action of chondroitin AC lyase Afrom Arthrobacter aurescens on both chondroitin-4- and -6-sulphateswas demonstrated viscosimetrically and confirmed using bothgradient PAGE and gel permeation chromatography. ChondroitinAC lyase F from Flavobacterium heparinum (Cytophagia heparinia)acted endolytically on the same substrates. Chondroitin B lyase(chondroitinase B, no EC number) from F.heparinum acted endolyticallyon dermatan sulphate giving a nearly identical action patternas observed for chondroitin ABC lyase acting on dermatan sulphate. action pattern chondroitin lyase glycosaminoglycan heparin lyase.  相似文献   

10.
The acid glycosaminoglycans were extracted from the skins of young rats less than 1 day post partum. The isolated products were fractionated by a cetylpyridinium chloride-cellulose column technique and identified by chemical analysis, electrophoretic mobility and susceptibility to testicular hyaluronidase digestion. Hyaluronic acid (56%) dermatan sulphate (15.6%) and chondroitin 6-sulphate (9.1%) were the major components, but chondroitin 4-sulphate, heparan sulphate and heparin were also present, together with two further fractions tentatively suggested to be a heparan sulphate-like fraction and a dermatan sulphate fraction, both of short chain length or low degree of sulphation.  相似文献   

11.
A sulphotransferase preparation from hen's uterus catalysed the transfer of sulphate from adenosine 3′-phosphate 5′-sulphatophosphate to N-desulphated heparan sulphate, heparan sulphate, N-desulphated heparin and dermatan sulphate. Heparin, chondroitin sulphate and hyaluronic acid were inactive as substrates for the enzyme. N-desulphated heparin was a much poorer substrate for the enzyme than N-desulphated heparan sulphate suggesting that properties of the substrate other than available glucosaminyl residues influenced enzyme activity. N-acetylation of N-desulphated heparin and N-desulphated heparan sulphate reduced their sulphate acceptor properties so it was unlikely that the N-acetyl groups of heparan sulphate facilitated its sulphatiion. Direct evidence for the transfer of [35S]sulphate to amino groups of N-desulphated haparan sulphate was obtained by subsequent isolation of glucosamine N-[35S]sulphate from heparan [35S]sulphate product. This was made possible through the use of a flavobacterial enzyme preparation which contained “heparitinase” activity but had been essentially freed of sulphatases. Attempts to transfer [35S]sulphate to glucosamine or N-acetylglucosamine were unsuccessfull.  相似文献   

12.
A heparin-binding protein was isolated from bovine uteri and purified to homogeneity. This protein appears as a double band of approx. 78 kDa in SDS/polyacrylamide-gel electrophoresis and has an isoelectric point of 5.2. The binding of heparin to this protein is saturable. No other glycosaminoglycan from mammalian tissue, such as hyaluronic acid, chondroitin sulphate, dermatan sulphate or keratan sulphate, binds to the 78 kDa protein. Dextran sulphate binds in a non-saturable fashion. Certain heparan sulphate polysaccharide structures are required for binding to the 78 kDa protein. Some proteoheparan sulphates, such as endothelial cell-surface proteoheparan sulphate, show only weak interaction with the 78 kDa protein in contrast with a basement-membrane proteoheparan sulphate from HR-9 cells. Antibodies against the 78 kDa protein inhibit binding of proteoheparan [35S]sulphate from basement membranes to smooth-muscle cells. Conventional antibodies, Fab fragments and some monoclonal antibodies, inhibit smooth-muscle cell proliferation in a similar range as that observed for heparin. The protein was detected in a variety of tissues and cells but not in blood cells. A possible role of this protein as a receptor for heparin or heparan sulphate and its function in the control of the arterial wall structure are discussed.  相似文献   

13.
1. N-Acetylgalactosamine 6-sulphate sulphatase was purified about 20000-fold from the soluble extract of human placenta with N-acetylgalactosamine 6-sulphate-glucuronic acid-N-acetyl[1-(3)H]galactosaminitol 6-sulphate as substrate in the activity assay. The enzyme appears to be a glycoprotein with a mol.wt. of about 100000 as determined by gel filtration. On gel electrophoresis in the presence of sodium dodecyl sulphate the major protein band had a mol.wt. of 78000. Variable charge heterogeneity was observed in several enzyme preparations. 2. The purified enzyme released up to one sulphate molecule from the disulphated trisaccharide. It was active towards N-acetylgalactosamine 6-sulphate and exhibited no measurable N-acetylglucosamine 6-sulphate sulphatase or any other known lysosomal sulphatase activity. Hydrolysis of [1-(3)H]galactitol 6-sulphate was achieved by incubation neither with a crude nor with a purified enzyme preparation. Chondroitin 6-sulphate and keratan sulphate, as well as heparin and heparan sulphate, served as competitive inhibitors of the enzyme. 3. Purified N-acetylgalactosamine 6-sulphate sulphatase activity was optimal at pH4.9 and 4.4 when assayed in 0.02m-sodium acetate buffer and at pH4.2 and 5.2 in 0.1m-sodium acetate buffer. A single pH-optimum at pH4.8 was observed for the crude enzyme and for the purified enzyme after mild periodate treatment. The sulphatase activity was inhibited by a variety of anions and cations and activated by thiol-specific and thiol reagents.  相似文献   

14.
A heparan sulphate sulphotransferase was partially purified from an ox lung homogenate by (NH(4))(2)SO(4) precipitation. Various glycosaminoglycans were assayed as sulphate acceptors with this enzyme. The highest acceptor activity was obtained with desulphated heparin and heparan sulphate, which indicates that sulphate transfer may be to free amino groups of the substrate. Some heparan sulphate was (35)S-labelled by incubation with the enzyme and re-isolated. On treatment of this heparan [(35)S]sulphate with nitrous acid and separation of the degradation products on Sephadex G-15, a major peak of radioactivity was obtained, and identified as [(35)S]sulphate by high-voltage electrophoresis at pH5.3. The [(35)S]sulphate is believed to be derived from N-[(35)S]sulphated groups of heparan [(35)S]-sulphate. That the ox lung preparation contained an N-sulphotransferase was confirmed by the isolation of 2-deoxy-2-[(35)S]sulphoamino-d-glucose as the major product from the flavobacterial degradation of heparan [(35)S]sulphate.  相似文献   

15.
We have previously shown that heparin is a potent inhibitor of a mammalian DNA topoisomerase I. We have now investigated the mechanism of its inhibition. This was carried out first by scrutinizing the structural features of heparin molecules responsible for the inhibition. Commercial heparin preparation was fractionated by antithrombin III-Sepharose into non-adsorbed, low-affinity and high-affinity fractions, of which only the high-affinity fraction of heparin is known to contain a specific oligosaccharide sequence responsible for the binding to antithrombin III. These fractions all exhibited essentially similar inhibitory activities. Furthermore, when chemically sulphated to an extent comparable with or higher than heparin, otherwise inactive glycosaminoglycans such as heparan sulphate, chondroitin 4-sulphate, dermatan sulphate and neutral polysaccharides such as dextran and amylose were converted into potent inhibitors. Sulphated dermatan sulphate, one of the model compounds, was further shown to bind competitively to the same sites on the enzyme as heparin. These observations strongly suggested that topoisomerase inhibition by heparin is attributable primarily, if not entirely, to the highly sulphated polyanionic nature of the molecules. In a second series of experiments we examined whether heparin inhibits only one or both of the topoisomerase reactions, i.e. nicking and re-joining. It was demonstrated that both reactions were inhibited by heparin, but the nicking reaction was more severely affected than was the re-joining reaction.  相似文献   

16.
Ordered conformations from the sodium salts of chondroitin 4-sulphate, dermatan sulphate and heparan sulphate were observed by X-ray diffraction. Chondroitin 4-sulphate shows similar threefold helical character to that previously reported for chondroitin 6-sulphate and hyaluronates. Dermatan sulphate forms an eightfold helix with an axial rise per disaccharide of 0.93nm, which favours the l-iduronic acid moiety in the normal C1 chair form. The layer-line spacing and axial projection in heparan sulphate of 1.86nm favours a tetrasaccharide repeat with glycosidic linkages alternating beta-d-(1-->4) and alpha-d-(1-->4).  相似文献   

17.
Various glycosaminoglycans have been subjected to affinity chromatography on immobilized bovine thrombin. Chondroitin sulphate, dermatan sulphate and heparan sulphate variants with a sulphate-to-hexosamine molar ratio of ~ 1 exhibited weak affinities. Heparan sulphate/heparin fractions of higher sulphate content could be separated into material with high and low affinity for thrombin. Removal of N-sulphate followed by N-acetylation did not affect binding, whereas oxidation and cleavage of non-sulphated hexuronate abolished the interaction. Heparan-related molecules of high thrombin-affinity comprised sequences where large blocks of sulphated iduronate-containing repeats were joined via a few repeats carrying non-sulphated iduronate or glucuronate to form continuous segments that were larger than decasaccharide.  相似文献   

18.
When cardiac muscle cells from mature rats were incubated in vitro in the presence of heparin (8.7 nmole ml-1) lipoprotein lipase activity appeared in the incubation medium. The intracellular activity of the enzyme remained unchanged. Other glycosaminoglycans (heparan sulphate, dermatan sulphate, keratan sulphate and chrondroitin 6-sulphate) at the same or higher concentrations were totally ineffective in producing any enzyme redistribution between cells and medium. The release seen in the presence of heparin was blocked by the presence of cycloheximide. Cycloheximide by contrast had no effect on the release observed in the presence of dexamethasone, The action of endogenous glycosaminoglycans are unlikely therefore to have a significant role to play in the movement of lipoprotein lipase in heart tissue in vivo.  相似文献   

19.
Glycosaminoglycans including dermatan sulphate, hyaluronan, heparan sulphate and heparin were chemically modified by O-sulphonation. By altering the reaction conditions, products having a different degree of O-sulphonation could be obtained. Glycosaminoglycan derivatives were prepared having no free hydroxyl groups, with sulphoester group/disaccharide unit ratios of 4.0 for dermatan sulphate and hyaluronan, and sulphoester and sulphamide group/disaccharide unit ratios of 4.22 and 4.88 for heparan sulphate and heparin, respectively. 1H NMR spectroscopy showed that the fully O-sulphonated hyaluronan derivative had a glucuronate residue with an altered conformation. Since glycosaminiglycans and their derivatives are often used as anticoagulant/antithrombotic agents, their anti-amidolytic activities were determined. The anti-factor IIa activity of fully O-sulphonated dermatan sulphate, hyaluronan and heparan sulphate ranged from 40 to 80 units/mg, while no anti-factor Xa activity of the fully O-sulphonated glycosaminoglycans was detected. These values are lower than those reported for low-molecular-weight heparins and are consistent with the requirement of an antithrombin III pentasaccharide binding site for anti-factor Xa activity. Interestingly, the anti-factor Xa of heparin is lost by chemical O-sulphonation.  相似文献   

20.
Structural requirements for heparan sulphate self-association   总被引:4,自引:0,他引:4  
To investigate heparan sulphate self-association, various sub-fractions of beef-lung heparan sulphate have been subjected to affinity chromatography on heparan sulphate-agarose. A particular variant of heparan sulphate was chiefly bound to matrices substituted with the same or cognate heparan sulphates. N-desulphation and N-acetylation abolished the chain-chain interaction. Also, dermatan sulphates and chondroitin sulphates showed affinity for heparan sulphate-agarose. [3H]Heparan sulphates that were bound to a heparan sulphate-agarose were desorbed by elution with the corresponding heparan sulphate chains and also with unrelated heparan sulphates, heparin, and the galactosaminoglycans to various degrees. However, the corresponding heparan sulphate species was the most efficient at low concentrations. Dextran sulphate was unable to desorb bound heparan sulphate. When the corresponding heparan sulphate was N-desulphated/N-acetylated, carboxyl-reduced, or periodate-oxidised (D-glucuronate), the modified polymer was unable to displace [3H]heparan sulphate from heparan sulphate-agarose. The displacing ability of heparin was also destroyed by periodate oxidation. It is concluded that self-interaction between heparan sulphate chains is strongly dependent on the overall molecular conformation. The N-sulphate and carboxylate groups as well as the integrity of the D-glucuronate residue are all essential for maintaining the proper secondary structure.  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