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.
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   

2.
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   

3.
Heparin was previously reported to potentiate the mitogenic activity of endothelial cell mitogens in a crude extract of bovine hypothalami (Thornton, S. C., Mueller, S. N., and Levine, E. M. (1983) Science 222, 623-625). We and others (Gospodarowicz, D., and Cheng, J. (1986) J. Cell. Physiol. 128, 475-484) have reported that the growth stimulatory effects of acidic fibroblast growth factor (aFGF) are potentiated in a similar manner. We have used these observations as the basis of an assay to characterize the importance of size, sulfation, and anticoagulant activity of heparin in mediating this effect. Partial nitrous acid depolymerization of heparin from porcine intestinal mucosa resulted in a mixture of heparin fragments, containing oligosaccharides ranging from disaccharides to polysaccharides of about 40 monosaccharides in length. This mixture was fractionated by ion exchange chromatography and gel permeation chromatography to obtain size-homogeneous oligosaccharides with different degrees of sulfation. Assay of these heparin-derived saccharides in the presence of a suboptimal concentration of aFGF revealed that a minimum chain length and a certain degree of sulfation is required in order to potentiate the action of aFGF. Low sulfate oligosaccharides (4-16 units) were unable to potentiate aFGF, whereas medium sulfate fractions of octadecasaccharides and larger were able to moderately potentiate aFGF. The potentiation of aFGF by the high sulfate fraction correlated with the saccharide size: 12 or more monosaccharide units were necessary to achieve potentiation equivalent to whole heparin, octa- and decasaccharides were mildly stimulatory, and hexasaccharides were without effect. In the absence of aFGF, intact heparin as well as all the oligosaccharides examined, inhibited the proliferation of capillary endothelial cells to approximately the same degree, between 20 and 50% inhibition. When a tetradecasaccharide was separated into a binding and a nonbinding fraction on matrix-bound antithrombin III, no difference was seen for these fractions in the endothelial cell proliferation assay. These results indicate that both size and sulfation of a heparin-derived oligosaccharide contribute to its ability to interact with aFGF and/or endothelial cells and that this interaction is independent of anticoagulant activity. In addition, our findings suggest that the inhibitory and potentiating effects of heparin on capillary endothelial cells have different structural requirements.  相似文献   

4.
Heparan sulphate by-products from the commercial manufacture of pig mucosal heparin were freed of chondroitin sulphate and fractionated according to anionic density. The fractions were treated with HNO2 at pH 1.5, and the resulting mixtures of oligosaccharides were reduced with NaB3H4 and analysed for their disaccharide composition by paper chromatography and by high-pressure liquid chromatography. The results show that the molar ratio of 2-O-sulpho-alpha-L-iduronosylanhydromannose to 6-O-sulpho-(2-O-sulpho-alpha-L-iduronosyl)anhydromannose decreased from 2.5 to 0.04 as the degree of sulphation of the fractions increased. In contrast, the molar ratio of 6-O-sulpho-(beta-D-glucuronosyl)anhydromannose to 6-O-sulpho-(alpha-L-iduronosyl)anhydromannose was approx. 2.4 in all heparan sulphate fractions and decreased to only half of this value in the most highly sulphated heparin fractions. These results are consistent with biosynthetic studies, which have shown that the N-sulpho-(2-O-sulpho-alpha-L-iduronosyl)D-glucosamine disaccharide is the metabolic precursor of the NO-disulpho-(2-O-sulpho-alpha-L-iduronosyl)-D-glucosamine disaccharide in heparin biosynthesis. The high-pressure liquid chromatography of the heparan sulphate oligosaccharides also revealed a number of unidentified oligosaccharides in the deamination mixtures.  相似文献   

5.
C(18) and C(8) bonded stationary phases dynamically coated with cetyltrimethylammonium (CTA) and strong anion exchange (SAX) were developed to obtain separations of oligosaccharide mixtures resulting from chemical or enzymatic depolymerization of heparin. With this method, the retention of sulfated oligosaccharides is directly adjustable depending on the amount of CTA adsorbed into the column. Oligosaccharides containing up to 20 sulfates were separated with a resolving power superior to that of conventional SAX analysis. The stability of the column coating enables hundreds of injections. Using ammonium methane sulfonate aqueous solutions as ultraviolet transparent mobile phases, it was possible to set up double detection, including selective detection of acetylated oligosaccharides. Analytical gel permeation chromatography was directly coupled to CTA-SAX, obtaining a two-dimensional profile of analyzed oligosaccharidic mixtures. A sequencing method of heparin oligosaccharides using partial depolymerization by heparinases according to their size and sulfation pattern and digest analysis by CTA-SAX was developed. A direct application of this method to the analysis of oligosaccharide mixtures obtained by complete digestion of heparins by heparinases I, II, and III was done. It allowed a reliable quantification of heparin building blocks. We also focused our attention on di- and tetrasaccharidic species containing the 3-O-sulfated glucosamines taken as markers of the active sites for antithrombin III. The method was also applied to more complex mixtures resulting from porcine heparin partially depolymerized with heparinase I. The specificity of the reaction was studied up to decasaccharidic fractions.  相似文献   

6.
An ion-pairing high-pressure liquid chromatography procedure was developed for analysis of mixtures of oligosaccharides generated by nitrous acid cleavage of heparin. Oligosaccharides were eluted from a Hi-Chrom 5S ODS (C18) column using mixtures of acetonitrile and buffers containing 40 mM ammonium phosphate and 1 mM tetrabutylammonium phosphate. Isocratic conditions were developed for optimal separation of a number of individual disaccharides and tetrasaccharides that were characterized previously (M.J. Bienkowski and H.E. Conrad (1985) J. Biol. Chem. 260, 356-365). These isocratic conditions were then coupled to obtain gradient elution conditions for the ion-pairing separations of mixtures of disaccharides and mixtures of tetrasaccharides. A comparison of the elution profiles obtained in the ion-pairing chromatography procedure with profiles obtained by anion-exchange high-pressure liquid chromatography profiles showed markedly better overall resolution by the ion-pairing procedure. As a result of this improved resolution, the new procedure showed the presence of previously unidentified products in the heparin oligosaccharide mixtures.  相似文献   

7.
Ion-pairing reversed-phased chromatography/mass spectrometry of heparin   总被引:1,自引:0,他引:1  
Heparin and heparin-derived components are widely applied anticoagulant drugs used for amongst other applications medical treatment of deep vein thrombosis and pulmonary embolism. Depolymerisation of native heparin results in complex mixtures of sulfated linear oligosaccharides that are usually not well characterised. In order to further characterise such mixtures, two on-line ion-pairing reverse-phased chromatography electrospray ionisation (ESI) mass spectrometry methods have been developed. One of the systems allows the determination of more than 200 components in a medium molecular weight heparin preparation, whereas the other system can be used to separate isomeric heparin oligosaccharides after previous separation according to size. This latter system allows semi-preparative isolation of isomeric heparin oligosaccharides. The experimental setup includes on-line cation exchange in order to prevent the ion-pairing reagent from entering the mass spectrometer.  相似文献   

8.
Chondroitin sulfate (CS) is a glycosaminoglycan consisting of repeating uronic acid, N-acetylgalactosamine disaccharide units {[HexAbeta/alpha(1-3)GalNAcbeta(1-4)](n)()}. CS chains are polydisperse with respect to chain length, sulfate content, and glucuronic acid epimerization content, resulting in a distribution of glycoforms for a chain bound to any given serine residue. Usually, CS glycoforms exist, differing in sulfation position and uronic acid epimerization. This work introduces a novel LC-MS/MS platform for the quantification of mixtures of CS oligosaccharides. The CS polysaccharides were partially depolymerized and labeled with either the light (d(0)) or heavy (d(4)) form of 2-anthranilic acid (2-AA). Excess reagent was removed, and mixtures of the CS standard (d(0)) and unknown (d(4)) were made. The CS mixture was subjected to size exclusion chromatography (SEC) with on-line electrospray ionization mass spectrometric detection in the negative ion mode. Tandem mass spectra were acquired, and quantification of unknown samples within the mixture was made using relative ion abundances of specific diagnostic ions. The high accuracy and precision of the glycomics platform were demonstrated using glycoform mixtures made from standard CS preparations. The CS glycoform analysis method was then applied to cartilage extract, versican, and several dermatan sulfate preparations. This work presents the first application of a glycomics platform for the quantification of CS oligosaccharide mixtures for obtaining specific information about the positions of GalNAc sulfation and uronic acid epimerization.  相似文献   

9.
Glycosaminoglycans have been implicated in the binding and activation of a variety of growth factors, cytokines, and chemokines. In this way, glycosaminoglycans are thought to participate in events such as development and wound repair. In particular, heparin and heparan sulfate have been well studied, and specific aspects of their structure dictate their participation in a variety of activities. In contrast, although dermatan sulfate participates in many of the same biological processes as heparin and heparan sulfate, the interactions of dermatan sulfate have been less well studied. Dermatan sulfate is abundant in the wound environment and binds and activates growth factors such as fibroblast growth factor-2 (FGF-2) and FGF-7, which are present during the wound repair process. To determine the minimum size and sulfation content of active dermatan sulfate oligosaccharides, dermatan sulfate was first digested and then separated by size exclusion high pressure liquid chromatography, and the activity to facilitate FGF-2 and FGF-7 was assayed by the cellular proliferation of cell lines expressing FGFR1 or FGFR2 IIIb. The minimum size required for the activation of FGF-2 was an octasaccharide and for FGF-7 a decasaccharide. Active fractions were rich in monosulfated, primarily 4-O-sulfated, disaccharides and iduronic acid. Increasing the sulfation to primarily 2/4-O-sulfated and 2/6-O-sulfated disaccharides did not increase activity. Cell proliferation decreased or was abolished with higher sulfated dermatan sulfate preparations. This indicated a preference for specific dermatan sulfate oligosaccharides capable of promoting FGF-2- and FGF-7-dependent cell proliferation. These data identify critical oligosaccharides that promote specific members of the FGF family that are important for wound repair and angiogenesis.  相似文献   

10.
Liquid chromatography/mass spectrometry (LC/MS) is applied to the analysis of complex mixtures of oligosaccharides obtained through the controlled, heparinase-catalyzed depolymerization of heparin. Reversed-phase ion-pairing chromatography, utilizing a volatile mobile phase, results in the high resolution separation of highly sulfated, heparin-derived oligosaccharides. Simultaneous detection by UV absorbance and electrospray ionization-mass spectrometry (ESI-MS) provides important structural information on the oligosaccharide components of this mixture. Highly sensitive and easily interpretable spectra were obtained through post-column addition of tributylamine in acetonitrile. High resolution mass spectrometry afforded elemental composition of many known and previously unknown heparin-derived oligosaccharides. UV in combination with MS detection led to the identification of oligosaccharides arising from the original non-reducing end (NRE) of the heparin chain. The structural identification of these oligosaccharides provided sequence from a reading frame that begins at the non-reducing terminus of the heparin chain. Interestingly, 16 NRE oligosaccharides are observed, having both an even and an odd number of saccharide residues, most of which are not predicted based on biosynthesis or known pathways of heparin catabolism. Quantification of these NRE oligosaccharides afforded a number-averaged molecular weight consistent with that expected for the pharmaceutical heparin used in this analysis. Molecular ions could be assigned for oligosaccharides as large as a tetradecasaccharide, having a mass of 4625 Da and a net charge of -32. Furthermore, MS detection was demonstrated for oligosaccharides with up to 30 saccharide units having a mass of >10000 Da and a net charge of -60.  相似文献   

11.
In the structural analysis of heparin and heparan sulfate, it is customary to combine or pool like-sized fractions obtained by size-exclusion chromatography (SEC) of enzymatically derived heparin oligosaccharides. In this study, we examine the heterogeneity of preparative-scale SEC fractions obtained from enzymatic digests of porcine intestinal mucosa heparin. Each fraction was profiled by capillary electrophoresis with UV detection (CE−UV) using a 60 mM formic acid running buffer at pH 3.43. Differences in the composition and relative concentration of components of the SEC fractions were observed for disaccharides and larger oligosaccharides. The heterogeneity of the fractions becomes more pronounced when heparin is digested using a heparin lyase cocktail. The heterogeneity of preparative SEC fractions was further investigated by reversed-phase ion-pairing ultraperformance liquid chromatography coupled with mass spectrometry (RPIP−UPLC−MS) using the ion-pairing reagent, tributylamine (Bu3N). Our results suggest that preliminary profiling of preparative SEC fractions prior to pooling may simplify efforts to identify and/or isolate rare structures.  相似文献   

12.
Heparan sulfate proteoglycans were extracted from rat brain microsomal membranes or whole forebrain with deoxycholate and purified from accompanying chondroitin sulfate proteoglycans and membrane glycoproteins by ion-exchange chromatography, affinity chromatography on lipoprotein lipase-Sepharose, and gel filtration. The proteoglycan has a molecular size of approximately 220,000, containing glycosaminoglycan chains of Mr = 14,000-15,000. In [3H]glucosamine-labeled heparan sulfate proteoglycans, approximately 22% of the radioactivity is present in glycoprotein oligosaccharides, consisting predominantly of N-glycosidically linked tri- and tetraantennary complex oligosaccharides (60%, some of which are sulfated) and O-glycosidic oligosaccharides (33%). Small amounts of chondroitin sulfate (4-6% of the total glycosaminoglycans) copurified with the heparan sulfate proteoglycan through a variety of fractionation procedures. Incubation of [35S]sulfate-labeled microsomes with heparin or 2 M NaCl released approximately 21 and 13%, respectively, of the total heparan sulfate, as compared to the 8-9% released by buffered saline or chondroitin sulfate and the 82% which is extracted by 0.2% deoxycholate. It therefore appears that there are at least two distinct types of association of heparan sulfate proteoglycans with brain membranes.  相似文献   

13.
Oligosaccharides were isolated from heparin and heparan sulfate by a procedure consisting of three major steps: (a) acid hydrolysis; (b) gel chromatography; and (c) cation exchange chromatography on an amino acid analyzer. To date, six new oligosaccharides have been isolated by this procedure and have been sequenced by a combination of NaB3H4-labeling and deaminative cleavage with nitrous acid. The structures of these oligosaccharides were as follows: 1. GlcN-GlcUA-GlcN 2. GlcN-IdUA-GlcN 3. GlcN-GlcUA-GlcN-GlcUA-GlcN 4. GlcN-IdUA-GlcN-GlcUA-GlcN 5. GlcN-GlcUA-GlcN-IdUA-GlcN 6. GlcN-IdUA-GlcN-IdUA-GlcN The linkage positions and anomeric configurations were assumed to be the same as in the polysaccharides from which the oligosaccharides originated. The usefulness of some of these oligosaccharides as enzyme substrates was tested after appropriate modifications and radioactive labeling. Oligosaccharides 2 and 3 were N-[35S]sulfated and were found to serve as substrates for heparan N-sulfate sulfatase (heparin sulfamidase), with a homogenate of cultured skin fibroblasts as enzyme source. Similarly, reduction of oligosaccharide 2 with NaB3H4 yielded a substrate for acetyl-CoA:alpha-D-glucosaminide N-acetyltransferase. Finally, the previously known disaccharide, 4-O-alpha-D-glucosaminyl-L-iduronic acid, which was isolated in the course of this work, was N-acetylated with [3H] acetic anhydride and was shown to be a substrate for N-acetyl-alpha-D-glucosaminidase.  相似文献   

14.
Monoclonal antibodies were raised against a conjugate between heparin oligosaccharides and human serum albumin. The oligosaccharides were prepared by partial nitrous acid degradation of heparin and were coupled to human serum albumin by reductive amination. Characterization of the antibodies secreted by one of the resulting clones showed that they recognize a determinant present in the oligosaccharide antigen, but not in intact heparin, nor in a variety of related polysaccharides. Degradation of heparin by nitrous acid generates a 2,5-anhydro-D-mannose residue at the reducing end of the resulting oligosaccharides, and it is concluded that this structure is essential for interaction with the antibodies. Reduced oligosaccharides (containing a terminal anhydromannitol residue) are also active. After gel chromatography of partially degraded heparin, the smallest components capable of binding to the antibodies were found in a tetrasaccharide fraction. Affinity chromatography on immobilized monoclonal antibodies separated this tetrasaccharide fraction into distinct populations of binding and nonbinding species. Structural analysis showed that the tetrasaccharide fraction that bound to the monoclonal antibodies contained one single component with the structure IdoA(2-OSO3)-GlcNSO3 (6-OSO3)-IdoA(2-OSO3)-aManR(6-OSO3), whereas the fraction that did not bind to the antibodies contained a mixture of different structures.  相似文献   

15.
Porcine mucosal heparin was partially depolymerized with heparinlyase I and then fractionated into low-molecularweight (<5000)and high-molecular-weight (>5000) oligosaccharides by pressurefiltration. The high-molecular-weight oligosaccharide mixture({small tilde}50 wt% of the starting heparin) also containedintact heparin. This intact polymer complicates oligosacsharidepurification. Thus, the low-molecular-weight fraction was usedto prepare homogeneous oligosaccharides for structural characterization.The low-molecular-weight oligosaccharide mixture was first fractionatedby low pressure gel permeation chromatography into size-uniformmixtures of disaccharides, tetrasaccharides, hexasaccharides,octasaccharides, decasaccharides, dodecasaccharides, tetradecasaccharidesand higher oligosaccharides. Each size-fractionated mixturewas then purified on the basis of charge by repetitive semi-preparativestrong-anion-exchange high-performance liquid chromatography.This approach has led to the isolation of 14 homogeneous oligosaccharidesfrom disaccharide to tetradecasaccharide. The purity of theseheparin-derived oligosaccharides was determined by gradientpolyacrylamide gel electrophoresis, analytical strong-anion-exchangehigh-performance liquid chromatography, capillary electrophoresisand one-dimensional nuclear resonance spectroscopy. The structureof these oligosaccharides was established using 600 MHz two-dimensionalnuclear resonance spectroscopy . The spectral methods used includedhomonuclear correlation spectroscopy, nuclear Overhauser effectspectroscopy and heteronuclear multiple quantum coherence spech-clscopy.The 1H/1H connectivities of the protons of each sugar residuein an oligosaccharide were established by two-dimensional homonuclearcorrelation spectroscopy, while 1H/13C assignments were madeusing 1H inverse detection. One- and two-dimensional nuclearresonance spectroscopic analysis of these heparin oligosaccharidesshowed two closely related groups of heparin-oligosaccharidesare afforded by enzymatic depolymerization of heparin. One groupis fully sulphated, having the structures  相似文献   

16.
Carbohydrate chip technology has a great potential for the high-throughput evaluation of carbohydrate-protein interactions. Herein, we report syntheses of novel sulfated oligosaccharides possessing heparin and heparan sulfate partial disaccharide structures, their immobilization on gold-coated chips to prepare array-type Sugar Chips, and evaluation of binding potencies of proteins by surface plasmon resonance (SPR) imaging technology. Sulfated oligosaccharides were efficiently synthesized from glucosamine and uronic acid moieties. Synthesized sulfated oligosaccharides were then easily immobilized on gold-coated chips using previously reported methods. The effectiveness of this analytical method was confirmed in binding experiments between the chips and heparin binding proteins, fibronectin and recombinant human von Willebrand factor A1 domain (rh-vWf-A1), where specific partial structures of heparin or heparan sulfate responsible for binding were identified.  相似文献   

17.
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   

18.
A simple panning procedure that allows for the evaluation of interactions between various heparin-like molecules and basic FGF has been developed. This assay measures the ability of compounds to inhibit the interaction of transfected human lymphoblastoid cells, UC 729-6 (UC cells), expressing hamster syndecan and basic FGF-coated plastic plates. The transfected cells bind rapidly to basic FGF-coated plates while the control cells do not bind well. Binding of the transfected cells to basic FGF was inhibited by heparin and heparin sulfate (HS), but not by chondroitin sulfate, dermatan sulfate, keratan sulfate, and hyaluronic acid. There was little inhibition of binding by chemically modified heparin such as completely desulfated, N-acetylated heparin, completely desulfated, N-sulfated heparin, and N-desulfated, N-acetylated heparin. These results suggested that both the N-sulfate and O-sulfate groups of heparin are required for binding to basic FGF. In addition, inhibition by oligosaccharides derived from depolymerized heparin increased with fragment size; partial inhibition was observed with oligosaccharides as small as hexamers. The biochemical basis for the binding of transfected cells to basic FGF was established by showing a significant increase of 35SO4 incorporation into HS. In particular, the level of 35SO4-HS in the trypsin-releasable (cell surface) pool increased fivefold. This increase was accounted for by demonstration of the presence of HS on immunoprecipitated syndecan from the transfected cells.  相似文献   

19.
We have developed methods for rapid fractionation of anionic oligosaccharides containing sulfate and/or sialic acid moieties by high-performance liquid chromatography (HPLC). Ion-exchange HPLC on amine-bearing columns (Micropak AX-10 and AX-5) at pH 4.0 is utilized to separate anionic oligosaccharides bearing zero, one, two, three, or four charges, independent of the identity of the amnionic moieties (sulfate and/or sialic acid). Ion-exchange HPLC at pH 1.7 allows separation of neutral, mono-, di-, and tetrasialylated, monosulfated, and disulfated oligosaccharides. Oligosaccharides containing three sialic acid residues and those bearing one each of sulfate and sialic acid, however, coelute at pH 1.7. Since the latter two oligosaccharide species separate at pH 4.0, analysis at pH 4.0 followed by analysis at pH 1.7 can be utilized to completely fractionate complex mixtures of sulfated and sialylated oligosaccharides. Ion-suppression amine adsorption HPLC has previously been shown to separate anionic oligosaccharides on the basis of net carbohydrate content (size). In this study we demonstrate the utility of ion-suppression amine adsorption HPLC for resolving sialylated oligosaccharide isomers which differ only in the linkages of sialic acid residues (alpha 2.3 vs alpha 2.6) and/or location of alpha 2,3- and alpha 2,6-linked sialic acid moieties on the peripheral branches of oligosaccharides. These two methods can be used in tandem to separate oligosaccharides, both analytically and preparatively, based on their number, types, and linkages of anionic moieties.  相似文献   

20.
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   

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