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.
Site-specific structural characterization of the glycosylation of human lecithin:cholesterol acyltransferase (LCAT) was carried out using microbore reversed-phase high performance liquid chromatography coupled with electrospray ionization mass spectrometry (HPLC/ESIMS). A recently described mass spectrometric technique involving monitoring of carbohydrate-specific fragment ions during HPLC/ESIMS was employed to locate eight different groups of glycopeptides in a digest of a human LCAT protein preparation. In addition to the four expected N-linked glycopeptides of LCAT, a di-O-linked glycopeptide was detected, as well as three additional glycopeptides. Structural information on the oligosaccharides from all eight glycopeptides was obtained by sequential glycosidase digestion of the glycopeptides followed by HPLC/ESIMS. All four potential N-linked glycosylation sites (Asn20, Asn84, Asn272, and Asn384) of LCAT were determined to contain sialylated triantennary and/or biantennary complex structures. Two unanticipated O-linked glycosylation sites were identified at Thr407 and Ser409 of the LCAT O-linked glycopeptide, each of which contain sialylated galactose beta 1-->3N-acetylgalactosamine structures. The three additional glycopeptides were determined to be from a copurifying protein, apolipoprotein D, which contains potential N-linked glycosylation sites at Asn45 and Asn78. These glycopeptides were determined to bear sialylated triantennary oligosaccharides or fucosylated sialylated biantennary oligosaccharides. Previous studies of LCAT indicated that removal of the glycosylation site at Asn272 converts this protein to a phospholipase (Francone OL, Evangelista L, Fielding CJ, 1993, Biochim Biophys Acta 1166:301-304). Our results indicate that the carbohydrate structures themselves are not the source of this functional discrimination; rather, it must be mediated by the structural environment around Asn272.  相似文献   

2.
We have carried out detailed structural studies of the glycopeptides of glycoprotein gD of herpes simplex virus types 1 and 2. We first examined and compared the number of N-asparagine-linked oligosaccharides present in each glycoprotein. We found that treatment of either pgD-1 or pgD-2 with endo-β-N-acetylglucosaminidase H (Endo H) generated three polypeptides which migrated more rapidly than pgD on gradient sodium dodecyl sulfate-polyacrylamide gels. Two of the faster-migrating polypeptides were labeled with [3H]mannose, suggesting that both pgD-1 and pgD-2 contained three N-asparagine-linked oligosaccharides. Second, we characterized the [3H]mannose-labeled tryptic peptides of pgD-1 and pgD-2. We found that both glycoproteins contained three tryptic glycopeptides, termed glycopeptides 1, 2, and 3. Gel filtration studies indicated that the molecular weights of these three peptides were approximately 10,000, 3,900, and 1,800, respectively, for both pgD-1 and pgD-2. Three methods were employed to determine the size of the attached oligosaccharides. First, the [3H]mannose-labeled glycopeptides were treated with Endo H, and the released oligosaccharide was chromatographed on Bio-Gel P6. The size of this molecule was estimated to be approximately 1,200 daltons. Second, Endo H treatment of [35S]methionine-labeled glycopeptide 2 reduced the molecular size of this peptide from approximately 3,900 to approximately 2,400 daltons. Third, glycopeptide 2 isolated from the gD-like molecule formed in the presence of tunicamycin was approximately 2,200 daltons. From these experiments, the size of each N-asparagine-linked oligosaccharide was estimated to be approximately 1,400 to 1,600 daltons. Our experiments indicated that glycopeptides 2 and 3 each contained one N-asparagine-linked oligosaccharide chain. Although glycopeptide 1 was large enough to accommodate more than one oligosaccharide chain, the experiments with Endo H treatment of the glycoprotein indicated that there were only three N-asparagine-linked oligosaccharides present in pgD-1 and pgD-2. Further studies of the tryptic glycopeptides by reverse-phase high-performance liquid chromatography indicated that all of the glycopeptides were hydrophobic in nature. In the case of glycopeptide 2, we observed that when the carbohydrate was not present, the hydrophobicity of the peptide increased. The properties of the tryptic glycopeptides of pgD-1 were compared with the properties predicted from the deduced amino acid sequence of gD-1. The size and amino acid composition compared favorably for glycopeptides 1 and 2. Glycopeptide 3 appeared to be somewhat smaller than would be predicted from the deduced sequence of gD-1. It appears that all three potential glycosylation sites predicted by the amino acid sequence are utilized in gD-1 and that a similar number of glycosylation sites are present in gD-2.  相似文献   

3.
The three tryptic glycopeptides from ovine lutropin, in which two were from the α-subunit (α-56 and α-82 glycopeptides) and one from the β-subunit (β-13 glycopeptide), have been isolated and their carbohydrate compositions analyzed. The results indicate that the α-56 glycopeptide has the highest amount of carbohydrate, whereas the β-13 glycopeptide has the least. In general, each of the glycopeptides has similar distribution of various sugars, i.e. high in mannose and glucosamine and low in fucose, sialic acid, galactose and galactosamine. Within the limit of experimental error, the sum of their carbohydrate composition is in agreement with the published data on the intact hormone or separated subunits.  相似文献   

4.
The site-specific characterization of the complex glycans in multiglycosylated proteins requires developing methods where the carbohydrates remain covalently bound to the protein. The complexity in the carbohydrate composition of α1-acid glycoprotein (AAG) makes it an ideal model protein for such development. AAG has five N-asparaginyl-linked glycosylation sites, each varying in its bi-, tri-, and tetraantennary glycan content. We present an on-line liquid chromatography/mass spectrometry (LC/MS) method that uses high-low cone voltage switching for in-source fragmentation to determine the structures of the complex glycans present on each site for the two gene products of AAG. High cone voltage caused carbohydrate fragmentation, leading to the generation of signature carbohydrate ions that we used as markers to identify the glycopeptides. Low cone voltage produced minimal carbohydrate fragmentation and enabled the identification and quantification of the intact oligosaccharide structures on each glycopeptide based on its monoisotopic mass and intensity. Quantitation was accomplished by using the intensities of peaks from deconvoluted and deisotoped mass spectra or from the areas of the extracted ion chromatograms from the tryptic peptide maps. The combined results from the two methods can be used to better characterize and quantitate site heterogeneity in multiglycosylated proteins.  相似文献   

5.
We previously reported on glycopeptide mapping of erythropoietin (EPO) by liquid chromatography/mass spectrometry (LC/MS). Using this method, glycopeptides in proteolytic digestion can be eluted before peptides, and are further separated on the basis of the carbohydrate structure. The detailed glycosylation at each glycosylation site can be elucidated based on mass chromatography and mass spectroscopy. In this study, we evaluated glycopeptide mapping with regard to its use in comparability assessment of glycoprotein products possessing multiple glycosylation sites. Models of closely related glycoprotein products used in this study are EPOs produced from three different sources. We previously reported that there are differences in the carbohydrate heterogeneity of these EPOs with regard to sialylation, acetylation, and sulphation patterns, using sugar mapping by LC/MS. In this paper, we demonstrated that glycopeptide mapping can distinguish site-specific glycosylation among these three EPOs and reveal the differences in acetylation, sialylation, and sulphation at each glycosylation site in one analysis. Our method can thus be useful in comparability assessment of therapeutic glycoproteins in terms of glycosylation.  相似文献   

6.
A method is presented for separation of tryptic glycopeptides-containing oligosaccharides of the N-asparagine-linked type. High performance liquid Chromatography (HPLC) of glycopeptides on a C18 reverse-phase system eluted with a gradient of 0%–50% acetonitrile in 0.1 M NaPO4 pH 2.2 resolves the two major glycosylation sites from the envelope glycoprotein (G) of vesicular stomatitis virus. Glycopeptides containing N-linked oligosaccharides of the complex type coelute with those containing N-linked oligosaccharides of the neutral, high mannose type, indicating that separation is based upon peptide rather than carbohydrate composition. The contribution of the carbohydrate component to glycopeptide elution, as determined by cleavage of the high mannose oligosaccharides with endo-β-Nacetylglucosaminidase H, is that of a significant, but minor, decrease in peptide retention time. Comparison of the tryptic glycopeptide profiles of G isolated from both wild type and mutant strains of VSV illustrates the rapid, reproducible, and quantitative nature of the technique. Through HPLC analysis of appropriately treated glycopeptides, it is possible to explore both the nature and extent of glycosylation at individual sites in glycoproteins in a single step.  相似文献   

7.
Proteomic techniques, such as HPLC coupled to tandem mass spectrometry (LC-MS/MS), have proved useful for the identification of specific glycosylation sites on glycoproteins (glycoproteomics). Glycosylation sites on glycopeptides produced by trypsinization of complex glycoprotein mixtures, however, are particularly difficult to identify both because a repertoire of glycans may be expressed at a particular glycosylation site, and because glycopeptides are usually present in relatively low abundance (2% to 5%) in peptide mixtures compared to nonglycosylated peptides. Previously reported methods to facilitate glycopeptide identification require either several pre-enrichment steps, involve complex derivatization procedures, or are restricted to a subset of all the glycan structures that are present in a glycoprotein mixture. Because the N-linked glycans expressed on tryptic glycopeptides contribute substantially to their mass, we demonstrate that size exclusion chromatography (SEC) provided a significant enrichment of N-linked glycopeptides relative to nonglycosylated peptides. The glycosylated peptides were then identified by LC-MS/MS after treatment with PNGase-F by the monoisotopic mass increase of 0.984 Da caused by the deglycosylation of the peptide. Analyses performed on human serum showed that this SEC glycopeptide isolation procedure results in at least a 3-fold increase in the total number of glycopeptides identified by LC-MS/MS, demonstrating that this simple, nonselective, rapid method is an effective tool to facilitate the identification of peptides with N-linked glycosylation sites.  相似文献   

8.
Electrophoretically homogeneous type 1 (GP-C1 and GP-C2), type 2 (GP-C3a and GP-C3b,) and type 3 (GP-D1, and GP-D2) glycopeptides fromAspergillus niger glucoamylase II (Manjunath and Raghavendra Rao, preceding paper) were separately treated with alkaline borohydride. The (\-eliminated oligosaccharides were subjected to single and sequential digestion with specific glycosidases and the products analysed by gas liquid chromatography. The studies revealed that carbohydrate moieties were present as mannose, Man-Man-, and trisaccharide structures, namely, (a) GIc-Man-Man-, (b) Gal-Man-Man, (c) Man-Man-Man-, (d) GlcNAc-Man-Man-, and (e) Xyl-Man-Man. None of the glycopeptides contained all the trisaccharide structures (a) to (e). Type 1 glycopeptide contained structures (a), (b) and (c); type 2, (a) and (d) and type 3, (a), (b) and (e). The number of carbohydrate units (mono-, di-and trisaccharides) present in the major glycopeptides was determined and tentative structures for the glycopeptides proposed. Carbohydrate units appeared to occur in clusters of 4 to 7 in each glycopeptide, a structure unique to the carbohydrate moiety inAspergillus niger glucoamylase. Based on carbohydrate analysis and yields of glycopeptide, the number of units of each type of glycopeptide present in glucoamylase II was tentatively calculated to give two of type Man:Glc:Gal = 12–15:l:l, one of type Man:Glc:GlcN = 10-l1:1:2 and one of type Man :GIc :Gal:Xyl = 4–8:0.1:0.5-0.8:0.3-1 glycopeptides.  相似文献   

9.
The characterization of site-specific glycosylation is traditionally dependent on the availability of suitable proteolytic cleavage sites between each glycosylated residue, so that peptides containing individual glycosylation sites are recovered. In the case of heavily glycosylated domains such as theO-glycosylated mucins, which have no available protease sites, this approach is not possible. Here we introduce a new method to gain site-specific compositional data on the oligosaccharides attached to a single amino acid. Using a model glycopeptide from a mutant human albumin Casebrook, glycosylated PTH-Asn was recovered after sequential solid-phase Edman degradation, subjected to acid hydrolysis and the sugars were identified by high performance anion exchange chromatography with pulsed amperometric detection. The PTH-Asn(Sac) derivative was further characterized by ionspray mass spectrometry. Comparison between an endoproteinase Glu-C glycopeptide and a tryptic glycopeptide showed that the oligosaccharide attached to Asn494 was stable after at least 10 cycles of Edman degradation.  相似文献   

10.
We describe here a two-dimensional mapping procedure which is capable of resolving glycopeptides isolated by lectin affinity chromatography from radioiodinated tryptic digests of glycoproteins. Glycopeptide maps were successfully produced for the model proteins alpha 1-acid glycoprotein and fetuin, as well as for the two surface glycoproteins gp90 and gp45 from equine infectious anemia virus (EIAV). Differences were detected in the glycopeptide maps obtained for the gp90 and gp45 components from two antigenically distinct strains of EIAV, demonstrating the ability of this procedure to detect variations in glycosylation in closely related glycoproteins. Thus this glycopeptide mapping technique provides a simple, rapid method to study changes in glycopeptides requiring only micrograms of glycoprotein.  相似文献   

11.
Prior to investigating the role of individual glycosylation sites in insulin receptor function, we are mapping the sites of glycosylation in the receptor. We report here a generally applicable methodology for the isolation and identification of glycosylation sites in cell surface glycoproteins. Human insulin receptors were labeled with [3H]-sugars using a CHO cell line transfected with the human receptor cDNA. Labelled receptors were mixed with receptors purified from human placental membranes and tryptic peptides prepared. Peptides were fractionated by gel filtration chromatography to limit the number of non-glycopeptides present. Peptides were then separated by reverse phase HPLC and glycopeptides identified by scintillation counting. Using this technique we have shown the insulin receptor to be glycosylated at Asn 397 and Asn 881. This increase the known number of occupied glycosylation sites to five.  相似文献   

12.
Capillary isoelectric focusing (CIEF) was used to profile the cellulase composition in complex fermentation samples of secreted proteins from Trichoderma reesei. The enzyme cellobiohydrolase I (CBH I, also referred to as Cel7A), a major component in these extracts, was purified from different strains and characterized using analytical methods such as CIEF, high-performance anion-exchange chromatography with pulsed amperometric detection (HPAEC–PAD), and capillary liquid chromatography–electrospray mass spectrometry (cLC–ESMS). ESMS was also used to monitor the extent of glycosylation in CBH I isolated from T. reesei strain RUT-C30 and two derivative mutant strains. Selective identification of tryptic N-linked glycopeptides was achieved using LC–ESMS on a quadrupole/time-of-flight instrument with a mixed scan function. The suspected glycopeptides were further analyzed by on-line tandem mass spectrometry to determine the nature of N-linked glycans and their attachment sites. This strategy enabled the identification of a high mannose glycan attached to Asn270 (predominantly Man8GlcNAc2) and single GlcNAc occupancy at Asn45 and Asn384 with some site heterogeneity depending on strains and fermentation conditions. The linker region of CBH I was shown to be extensively glycosylated with di-, and tri-saccharides at Thr and Ser residues as indicated by MALDI-TOF and HPAEC–PAD experiments. Additional heterogeneity was noted in the CBH I linker peptide of RUT-C30 strain with the presence of a phosphorylated di-saccharide.  相似文献   

13.
Glycan microarray technology has become a successful tool for studying protein–carbohydrate interactions, but a limitation has been the laborious synthesis of glycan structures by enzymatic and chemical methods. Here we describe a new method to generate quantifiable glycan libraries from natural sources by combining widely used protease digestion of glycoproteins and Fmoc chemistry. Glycoproteins including chicken ovalbumin, bovine fetuin, and horseradish peroxidase (HRP) were digested by Pronase, protected by FmocCl, and efficiently separated by 2D-HPLC. We show that glycans from HRP glycopeptides separated by HPLC and fluorescence monitoring retained their natural reducing end structures, mostly core α1,3-fucose and core α1,2-xylose. After simple Fmoc deprotection, the glycans were printed on NHS-activated glass slides. The glycans were interrogated using plant lectins and antibodies in sera from mice infected with Schistosoma mansoni, which revealed the presence of both IgM and IgG antibody responses to HRP glycopeptides. This simple approach to glycopeptide purification and conjugation allows for the development of natural glycopeptide microarrays without the need to remove and derivatize glycans and potentially compromise their reducing end determinants.  相似文献   

14.
Two LC-ESI-MS methods for the analysis of antibody glycosylation are presented. In the first approach, tryptic glycopeptides are separated by RP chromatography and analyzed by ESI-MS. This "glycopeptide strategy" allows a protein- and subclass-specific quantitation of both neutral and sialylated glycan structures. Additional information about under- or deglycosylation and the protein backbone, e.g., termini, can be extracted from the same data. In the second LC-ESI-MS method, released oligosaccharides are separated on porous graphitic carbon (PGC). A complete structural assignment of neutral and sialylated oligosaccharides occurring on antibodies is thereby achieved in one chromatographic run. The two methods were applied to polyclonal human IgG, to commercial mAb expressed in CHO cells (Rituximab, Xolair, and Herceptin), in SP2/0 (Erbitux and Remicade) or NS0 cells (Zenapax) and the anti-HIV antibody 4E10 produced either in CHO cells or in a human cell line. Both methods require comparably little sample preparation and can be applied to SDS-PAGE bands. They both outperform non-MS methods in terms of reliability of peak assignment and MALDI-MS of underivatized glycans with regard to the recording of sialylated structures. Regarding fast and yet detailed structural assignment, LC-MS on graphitic carbon supersedes all other current methods.  相似文献   

15.
Immunoglobulin A1 (IgA1) from normal human serum is known to have O-linked sugar chains, sialylated Galβ1,3GalNAc, in the hinge portion. In order to reduce the microheterogenity of the sugar chain, the hinge glycopeptide prepared from IgA1 was sequentially treated with neuraminidase and β-galactosidase. The asialo-, agalacto-hinge glycopeptide (HGP-SG) composed of a 33-mer peptide (HP33) and N-acetylgalactosamine (GalNAc) residues was obtained. The HGP-SG was separated into three major peaks, A, B and C, by high-performance liquid chromatography (HPLC). Each glycopeptide fraction was further separated by capillary electrophoresis (CE). Peaks A, B and C with HPLC abundantly contained HP33 bearing five and six N-acetylgalactosamine residues (HGP33-5,6GN), HGP33-4,5GN and HGP33-3,4GN, respectively. Among these glycopeptide peaks, only the HGP33-5GN peak was partly split into two peaks based on the CE analysis – HGP33-5GN-α and -β. The glycopeptide, HGP25-5GN shortened by the thermolysin digest of HGP33-SG was also well separated into the α and β forms by CE analysis. No differences in their mass and peptide portion were observed between HGP25-5GN-α and -β. Therefore, the obtained result might indicate that HGP25-5GN-α was an isomer of HGP25-5GN-β differing in its stereospecific structure of the peptide portion and/or the attachment site of the GalNAc residue.  相似文献   

16.
Human beta-hexosaminidase (EC 3.2.1.52) is a lysosomal enzyme that hydrolyzes terminal N-acetylhexosamines from GM2 ganglioside, oligosaccharides, and other carbohydrate-containing macromolecules. There are two major forms of hexosaminidase: hexosaminidase A, with the structure alpha(beta a beta b), and hexosaminidase B, 2(beta a beta b). Like other lysosomal proteins, hexosaminidase is targeted to its destination via glycosylation and processing in the rough endoplasmic reticulum and Golgi apparatus. Phosphorylation of specific mannose residues allows binding of the protein to the phosphomannosyl receptor and transfer to the lysosome. In order to define the structure and placement of the oligosaccharides in mature hexosaminidase and thus identify candidate mannose 6-phosphate recipient sites, the major tryptic/chymotryptic glycopeptides from each isozyme were purified by reverse-phase high-performance liquid chromatography. Two major concanavalin A binding glycopeptides, localized to the beta b chain, and one non concanavalin A binding glycopeptide, localized to the beta a chain, were found associated with the beta-subunit in both hexosaminidase A and hexosaminidase B. A single major concanavalin A binding glycopeptide was found to be associated with the alpha subunit of hexosaminidase A. The oligosaccharide structures were determined by nuclear magnetic resonance spectrometry. Two of them, the alpha and one of the beta b glycans, contained a Man3-GlcNAc2 structure, while the remaining one on the beta b chain was composed of a mixture of Man5-7-GlcNAc2 glycans. The unique glycopeptide associated with the beta a chain contained a single GlcNAc residue. Thus, all three mature polypeptides comprising the alpha and beta subunits of hexosaminidase contain carbohydrate, the structures of which have the appearance of being partially degraded in the lysosome. In the alpha chain we found only one possible site for in vivo phosphorylation. In the beta it is unclear if only one or all three of the sites could have contained phosphate. However, mature placental hexosaminidase A and B can be rephosphorylated in vitro. This requires the presence of an oligosaccharide containing an alpha 1,2-linked mannose residue. Only the single Man6-7 (of the Man5-7-GlcNAc2 glycans) containing site on the beta b chain retains this type of residue. Therefore, this site may act as the sole in vitro substrate in both of the mature isozymes for the phosphotransferase.  相似文献   

17.
Ceruloplasmin has ferroxidase activity and plays an essential role in iron metabolism. In this study, a site-specific glycosylation analysis of human ceruloplasmin (CP) was carried out using reversed-phase high-performance liquid chromatography with electrospray ionization tandem mass spectrometry (LC-ESI-MS/MS). A tryptic digest of carboxymethylated CP was subjected to LC-ESI-MS/MS. Product ion spectra acquired data-dependently were used for both distinction of the glycopeptides from the peptides using the carbohydrate B-ions, such as m/z 204 (HexNAc) and m/z 366 (HexHexNAc), and identification of the peptide moiety of the glycopeptide based on the presence of the b- and y-series ions derived from the peptide. Oligosaccharide composition was deduced from the molecular weight calculated from the observed mass of the glycopeptide and theoretical mass of the peptide. Of the seven potential N-glycosylation sites, four (Asn119, Asn339, Asn378, and Asn743) were occupied by a sialylated biantennary or triantennary oligosaccharide with fucose residues (0, 1, or 2). A small amount of sialylated tetraantennary oligosaccharide was detected. Exoglycosidase digestion suggested that fucose residues were linked to reducing end GlcNAc in biantennary oligosaccharides and to reducing end and/or alpha1-3 to outer arms GlcNAc in triantennary oligosaccharides and that roughly one of the antennas in triantennary oligosaccharides was alpha2-3 sialylated and occasionally alpha1-3 fucosylated at GlcNAc.  相似文献   

18.
Changes in the glycosylation of some serum proteins are associated with certain diseases. In this study, we performed simultaneous site-specific glycosylation analysis of abundant serum glycoproteins by LC/Qq-TOF MS of human serum tryptic digest, the albumin of which was depleted. The glycopeptide peaks on the chromatogram were basically assigned by database searching with modified peak-list text files of MS/MS spectra and then based on mass differences of glycan units from characterized glycopeptides. Glycopeptide of IgG, haptoglobin and ceruloplasmin were confirmed by means of a comparison of their retention times and m/z values with those obtained by LC/MS of commercially available glycoproteins. Mass spectrometric carbohydrate heterogeneity in the assigned glycopeptides was analyzed by an additional LC/MS. We successfully demonstrated site-specific glycosylation of 23 sites in abundant serum glycoproteins.  相似文献   

19.
A beta-N-acetylgalactosaminyltransferase that preferentially transferred N-acetylgalactosamine to Sd(a-) Tamm-Horsfall glycoprotein was found in guinea-pig kidney microsomal preparations. This enzyme was kidney-specific and was able to transfer the sugar to other glycoproteins, such as fetuin and alpha 1-acidic glycoprotein. The presence of sialic acid in the acceptors was essential for the transferase activity when either glycoproteins or their Pronase glycopeptides were used as acceptors. Two glycopeptides (Tamm-Horsfall glycopeptides I and II) with a different carbohydrate composition were separated by DEAE-Sephacel chromatography from Pronase-digested Tamm-Horsfall glycoprotein. The amount of N-acetylgalactosamine transferred to glycopeptides by the enzyme correlated with their degree of sialylation. Enzymic digestion of N-[14C]acetylgalactosamine-labelled Tamm-Horsfall glycopeptide II showed that the transferred sugar was susceptible to beta-N-hexosaminidase. The amount of sugar cleaved by beta-hexosaminidase was strongly increased when the labelled Tamm-Horsfall glycopeptide II was pretreated with mild acid hydrolysis, a procedure that removed the sialic acid residues. Alkaline borohydride treatment of the labelled Tamm-Horsfall glycopeptide II did not release radioactivity, thus indicating that enzymic glycosylation took place at the N-asparagine-linked oligosaccharide units of Tamm-Horsfall glycoprotein.  相似文献   

20.
The carbohydrate portion of the G glycoprotein of vesicular stomatitis virus (VSV) grown in CHO cells (CHO/VSV) has been fractionated on BioGelP6, concanavalin A-Sepharose, and pea lectin-agarose. The results suggest that, in addition to sialic acid and fucose heterogeneity, the asparagine-linked complex carbohydrate moieties of CHO/VSV also display branching heterogeneity. Although the majority of the glycopeptides bind to concanavalin A-Sepharose in a manner typical of certain biantennary carbohydrate structures, a significant proportion do not bind to the lectin. The latter behavior is typical of tri- or tetraantennary (branched) carbohydrate structures. The CHO/VSV glycopeptides which do not bind to concanavalin A-Sepharose separate into bound and unbound fractions on pea lectin-agarose suggesting that they include at least two different types of (branched) carbohydrate structures. Glycopeptides from the G glycoprotein of VSV grown in two, independently derived CHO glycosylation mutants which belong to complementation group 4 (Lec4 mutants) were examined in the same manner. In contrast to glycopeptides from CHO/VSV, glycopeptides from Lec4/VSV which passed through concanavalin A-Sepharose did not contain a component which subsequently bound to pea lectin-agarose. A glycopeptide fraction with these lectin-binding properties was also missing from cell surface glycopeptides derived from Lec4 cells. The combined results are consistent with the hypothesis that Lec4 CHO glycosylation mutants lack a glycosyltransferase activity responsible for the addition of a (branch) N-acetylglucosamine residue linked β1,6 to mannose.  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