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 62 毫秒
1.
Two CHO glycosylation mutants that were previously shown to lack N-linked carbohydrates with GlcNAc beta 1,6Man alpha 1,6 branches, and to belong to the same genetic complementation group, are shown here to differ in the activity of N-acetylglucosaminyltransferase V (GlcNAc-TV) (UDP-GlcNA: alpha 1,6mannose beta-N-acetylglucosaminyltransferase V). One mutant, Lec4, has no detectable GlcNAc-TV activity whereas the other, now termed Lec4A, has activity equivalent to that of parental CHO in detergent cell extracts. However, Lec4A GlcNAc-TV can be distinguished from CHO GlcNAc-TV on the basis of its increased sensitivity to heat inactivation and its altered subcellular compartmentalization. Sucrose density gradient fractionation shows that the major portion of GlcNAc-TV from Lec4A cells cofractionates with membranes of the ER instead of Golgi membranes where GlcNAc-TV is localized in parental CHO cells. Other experiments show that Lec4A GlcNAc-TV is not concentrated in lysosomes, or in a post-Golgi compartment, or at the cell surface. The altered localization in Lec4A cells is specific for GlcNAc-TV because two other Lec4A Golgi transferases cofractionate at the density of Golgi membranes. The combined data suggest that both lec4 and lec4A mutations affect the structural gene for GlcNAc-TV, causing either the loss of GlcNAc-TV activity (lec4) or its miscompartmentalization (lec4A). The identification of the Lec4A defect indicates that appropriate screening of different glycosylation-defective mutants should enable the isolation of other mammalian cell trafficking mutants.  相似文献   

2.
Carbohydrate moieties derived from the G glycoprotein of Vesicular Stomatitis Virus (VSV) grown in parental Chinese hamster ovary (CHO) cells and the glycosylation mutant Lec4 have been analyzed by high-field 1H NMR spectroscopy. The major glycopeptides of CHOVSV and Lec4VSV were purified by their ability to bind to concanavalin A-Sepharose. The carbohydrates in this fraction are of the biantennary, complex type with heterogeneity in the presence of α(2,3)-linked sialic acid and α(1,6)-linked fucose residues. A minor CHOVSV glycopeptide fraction, which does not bind to concanavalin A-Sepharose but which binds to pea lectin-agarose, was also investigated by 1H NMR spectroscopy. These carbohydrates are complex moieties which appear to contain N-acetylglucosamine in β(1,6) linkage. Their spectral properties are most similar to those of a triantennary complex oligosaccharide containing a 2,6-disubstituted mannose α(1,6) residue. Carbohydrates of this type are not found among the glycopeptides of VSV grown in the Lec4 CHO glycosylation mutant.  相似文献   

3.
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   

4.
A correlation between increased beta-1,6 branching of N-linked carbohydrates and the ability of a cell to metastasize or to form a tumor has been observed in several experimental models. Lec9 Chinese hamster ovary (CHO) mutants exhibit a drastic reduction in tumorigenicity in nude mice, and this phenotype directly correlates with their ability to attach an increased proportion of beta-1,6-branched carbohydrates to the G glycoprotein of vesicular stomatitis virus (J. Ripka, S. Shin, and P. Stanley, Mol. Cell. Biol. 6:1268-1275, 1986). In this paper we provide evidence that cellular carbohydrates from Lec9 cells also contain an increased proportion of beta-1,6-branched carbohydrates, although they do not possess significantly increased activity of the beta-1,6 branching enzyme (GlcNAc-transferase V). Biosynthetic labeling experiments show that a substantial degree of underglycosylation occurs in Lec9 cells and that this affects several classes of glycoproteins. Lec9 cells synthesize ca. 40-fold less Glc3Man9GlcNAc2-P-P-lipid and ca. 2-fold less Man5GlcNAc2-P-P-lipid than parental cells do. In addition, Lec9 cells possess ca. fivefold less protein-bound oligosaccharide intermediates, and one major species is resistant to release by endo-beta-N-acetylglucosaminidase H (endo H). Membranes of Lec9 cells exhibit normal mannosylphosphoryldolichol synthase, glucosylphosphoryldolichol synthase, and N-acetylglucosaminylphosphate transferase activities in the presence of exogenous dolichyl phosphate. However, in the absence of exogenous dolichyl phosphate, mannosylphosphoryldolichol synthase and glucosylphosphoryldolichol synthase activities are reduced in membranes of Lec9 cells, indicating that membranes of Lec9 cells are deficient in lipid phosphate. This was confirmed by analysis of lipids labeled by [3H]mevalonate, which showed that Lec9 cells have less lipid phosphate than parental CHO cells. Mechanisms by which a defect in the synthesis of dolichol-oligosaccharides might alter the degree of beta-1,6 branching in N-linked carbohydrates are discussed.  相似文献   

5.
Identifying biological roles for mammalian glycans and the pathways by which they are synthesized has been greatly facilitated by investigations of glycosylation mutants of cultured cell lines and model organisms. Chinese hamster ovary (CHO) glycosylation mutants isolated on the basis of their lectin resistance have been particularly useful for glycosylation engineering of recombinant glycoproteins. To further enhance the application of these mutants, and to obtain insights into the effects of altering one specific glycosyltransferase or glycosylation activity on the overall expression of cellular glycans, an analysis of the N-glycans and major O-glycans of a panel of CHO mutants was performed using glycomic analyses anchored by matrix-assisted laser desorption ionization-time of flight/time of flight mass spectrometry. We report here the complement of the major N-glycans and O-glycans present in nine distinct CHO glycosylation mutants. Parent CHO cells grown in monolayer versus suspension culture had similar profiles of N- and O-GalNAc glycans, although the profiles of glycosylation mutants Lec1, Lec2, Lec3.2.8.1, Lec4, LEC10, LEC11, LEC12, Lec13, and LEC30 were consistent with available genetic and biochemical data. However, the complexity of the range of N-glycans observed was unexpected. Several of the complex N-glycan profiles contained structures of m/z ∼13,000 representing complex N-glycans with a total of 26 N-acetyllactosamine (Galβ1–4GlcNAc)n units. Importantly, the LEC11, LEC12, and LEC30 CHO mutants exhibited unique complements of fucosylated complex N-glycans terminating in Lewisx and sialyl-Lewisx determinants. This analysis reveals the larger-than-expected complexity of N-glycans in CHO cell mutants that may be used in a broad variety of functional glycomics studies and for making recombinant glycoproteins.  相似文献   

6.
Glycolipid transport between compartments of the Golgi apparatus has been reconstituted in a cell free system. Transport of lactosylceramide (galactose beta 1-4-glucose-ceramide) was followed from a donor to an acceptor Golgi population. The major glycolipid in CHO cells is GM3 (sialic acid alpha 2-3 galactose beta 1-4-glucose-ceramide). Donor membranes were derived from a Chinese hamster ovary (CHO) cell mutant (Lec2) deficient in the Golgi CMP-sialic acid transporter, and therefore contained lactosylceramide as the predominant glycolipid. Acceptor Golgi apparatus was prepared from another mutant, Lec8, which is defective in UDP-Gal transport. Thus, glucosylceramide is the major glycolipid in Lec8 cells. Transport was measured by the incorporation of labeled sialic acid into lactosylceramide (present originally in the donor) by transport to acceptor membranes, forming GM3. This incorporation was dependent on ATP, cytosolic components, intact membranes, and elevated temperature. Donor membranes were prepared from Lec2 cells infected with vesicular stomatitus virus (VSV). These membranes therefore contain the VSV membrane glycoprotein, G protein. Donor membranes derived from VSV-infected cells could then be used to monitor both glycolipid and glycoprotein transport. Transport of these two types of molecules between Golgi compartments was compared biochemically and kinetically. Glycolipid transport required the N- ethylmaleimide sensitive factor previously shown to act in glycoprotein transport (Glick, B. S., and J. E. Rothman. 1987. Nature [Lond.]. 326:309-312; Rothman, J. E. 1987. J. Biol. Chem. 262:12502-12510). GTP gamma S inhibited glycolipid and glycoprotein transport similarly. The kinetics of transport of glycolipid and glycoprotein were also compared. The kinetics of transport to the end of the pathway were similar, as were the kinetics of movement into a defined transport intermediate. It is concluded that glycolipid and glycoprotein transport through the Golgi occur by similar if not identical mechanisms.  相似文献   

7.
Post-translational modifications (PTMs) of proteins play important roles in the physiology of eukaryotes. In the PTMs, non-reversible glycosylations are classified as N-glycosylations and O-glycosylations, and are catalyzed by various glycosidases and glycosyltransferases. However, β-glycosidases are not known to play a role in N- and O-glycan processing, although both glycans provide partial structures as substrates for β-galactosidase and β-N-acetylglucosaminidase in the Golgi apparatus of human cells. We explored human Golgi β-galactosidase using fluorescent substrates based on a quinone methide cleavage (QMC) substrate design platform that was previously developed to image exo-type glycosidases in living cells. As a result, we discovered a novel Golgi β-galactosidase in human cells. It is possible to predict a novel and important function in glycan processing of this β-galactosidase, because various β-galactosyl linkages in N- and O-glycans exist in Golgi apparatus. In addition, these results show that the QMC platform is excellent for imaging exo-type glycosidases.  相似文献   

8.
Nucleotide sugar transporters, encoded by the SLC35 gene family, deliver nucleotide sugars throughout the cell for various glycosyltransferase-catalyzed glycosylation reactions. GlcNAc, in the form of UDP-GlcNAc, and galactose, as UDP-Gal, are delivered into the Golgi apparatus by SLC35A3 and SLC35A2 transporters, respectively. However, although the UDP-Gal transporting activity of SLC35A2 has been clearly demonstrated, UDP-GlcNAc delivery by SLC35A3 is not fully understood. Therefore, we analyzed a panel of CHO, HEK293T, and HepG2 cell lines including WT cells, SLC35A2 knockouts, SLC35A3 knockouts, and double-knockout cells. Cells lacking SLC35A2 displayed significant changes in N- and O-glycan synthesis. However, in SLC35A3-knockout CHO cells, only limited changes were observed; GlcNAc was still incorporated into N-glycans, but complex type N-glycan branching was impaired, although UDP-GlcNAc transport into Golgi vesicles was not decreased. In SLC35A3-knockout HEK293T cells, UDP-GlcNAc transport was significantly decreased but not completely abolished. However, N-glycan branching was not impaired in these cells. In CHO and HEK293T cells, the effect of SLC35A3 deficiency on N-glycan branching was potentiated in the absence of SLC35A2. Moreover, in SLC35A3-knockout HEK293T and HepG2 cells, GlcNAc was still incorporated into O-glycans. However, in the case of HepG2 cells, no qualitative changes in N-glycans between WT and SLC35A3 knockout cells nor between SLC35A2 knockout and double-knockout cells were observed. These findings suggest that SLC35A3 may not be the primary UDP-GlcNAc transporter and/or different mechanisms of UDP-GlcNAc transport into the Golgi apparatus may exist.  相似文献   

9.
During studies on the fucosylation of endogenous proteins inparental (Pro5) and N-acetyl-D-glucosamine (GlcNAc) transferaseI-deficient (Lec1) Chinese hamster ovary (CHO) cells, we observedthat Lec1 cells incorporate  相似文献   

10.
Disruption of Golgi α-mannosidase II activity can result in type II congenital dyserythropoietic anemia and induce lupus-like autoimmunity in mice. Here, we isolated a mutant human embryonic kidney (HEK) 293T cell line called Lec36, which displays sensitivity to ricin that lies between the parental HEK 293T cells, in which the secreted and membrane-expressed proteins are dominated by complex-type glycosylation, and 293S Lec1 cells, which produce only oligomannose-type N-linked glycans. Stem cell marker 19A was transiently expressed in the HEK 293T Lec36 cells and in parental HEK 293T cells with and without the potent Golgi α-mannosidase II inhibitor, swainsonine. Negative ion nano-electrospray ionization mass spectra of the 19A N-linked glycans from HEK 293T Lec36 and swainsonine-treated HEK 293T cells were qualitatively indistinguishable and, as shown by collision-induced dissociation spectra, were dominated by hybrid-type glycosylation. Nucleotide sequencing revealed mutations in each allele of MAN2A1, the gene encoding Golgi α-mannosidase II: a point mutation that mapped to the active site was found in one allele, and an in-frame deletion of 12 nucleotides was found in the other allele. Expression of the wild type but not the mutant MAN2A1 alleles in Lec36 cells restored processing of the 19A reporter glycoprotein to complex-type glycosylation. The Lec36 cell line will be useful for expressing therapeutic glycoproteins with hybrid-type glycans and as a sensitive host for detecting mutations in human MAN2A1 causing type II congenital dyserythropoietic anemia.Mammalian N-linked glycosylation is characterized by significant chemical heterogeneity generated by an array of competing glycosidases and glycosyltransferases (1). The structural analysis of recombinant glycoproteins, such as human erythropoietin (2, 3), has illustrated the capacity of mammalian expression systems for generating diverse N-linked glycans.Heterogeneity develops during egress of a glycoprotein through the secretory system (1). N-linked glycosylation is initiated in the rough endoplasmic reticulum (ER)4 by the co-translational transfer of Glc3Man9GlcNAc2 to the asparagine residues of the glycosylation sequon. In the absence of protein misfolding, hydrolysis by ER α-mannosidase I plus α-glucosidase I and II results in the transfer of glycoproteins dominated by the D1,D3 isomer of Man8GlcNAc2 glycans to the Golgi apparatus (4). Further processing by Golgi α-mannosidases IA–C generates Man5GlcNAc2 (57), the principle substrate for UDP-N-acetyl-d-glucosamine:α-3-d-mannoside β1,2-N-acetylglucosaminyltransferase I (GnT I). The action of this enzyme yields classic hybrid-type glycans with mannosyl 6-antennae and processed 3-antennae (1). In the absence of the GnT III-mediated addition of bisecting GlcNAc, the two terminal α-mannose residues of the 6-antenna of hybrid-type glycans are cleaved by Golgi α-mannosidase II, forming mono-antennary complex-type glycans. These may then be processed by N-acetylglucosaminyltransferases, generating multiantennary complex-type glycans of enormous potential heterogeneity following the sequential transfer of monosaccharides such as galactose, N-acetylgalactosamine, fucose, and N-acetylneuraminic acid (8).The importance of this carbohydrate diversity in metazoan biology is illustrated by the disease phenotypes that manifest when the biosynthesis of particular glycoforms is disrupted. In humans, about 12 congenital disorders of glycosylation (CDG) have been identified with defects in the biosynthesis of N-linked glycans (9). One disorder characterized by changes in glycosylation is congenital dyserythropoietic anemia type II (hereditary erythroblastic multinuclearity with a positive acidified serum lysis test (HEMPAS)) (10, 11). HEMPAS is a heterogenous autosomal recessive disorder that renders erythrocytes prone to lysis. Although the precise molecular basis of HEMPAS remains to be determined, it is characterized by either a reduction in β1→4-galactosyltransferase, GnT II, or, in some patients, Golgi α-mannosidase II activity (11, 12). Interestingly, the increase in cell surface terminal mannose in mice deficient in Golgi α-mannosidase II leads to autoimmunity through chronic activation of the innate immune system (13, 14).Lectin-resistant (Lec) cell lines harboring loss- or gain-of-function mutations affecting the biosynthesis of N-glycans have emerged as powerful tools for the investigation of these disorders (15). For example, genetic complementation using Lec2, containing a mutation in the cytosine monophosphate sialic acid transporter, was used to identify a novel CDG, type IIf (16). Lectin-resistant cell lines can also be used as hosts to study naturally occurring mutations, as in the case of CHO Lec23 cells used to screen α-glucosidase I mutations in CDG, type IIb (17). Other applications of lectin-resistant cell lines include the expression of specific glycoforms of therapeutic glycoproteins. Manipulating the structure of their carbohydrate moieties modulates the pharmacological properties of glycoproteins by altering their bioactivity, serum half-life, and/or tissue tropism (18). For example, β-glucocerebrosidase expressed in CHO Lec1 cells (deficient in GnT I activity) exhibits mannosylation and improved macrophage uptake for the treatment of Gaucher disease (19). Lectin-resistant CHO cell lines have also been used to improve the crystallizability of glycoproteins for structural determination by x-ray crystallography (2024).The expression of therapeutic glycoproteins as one or more defined “glycoforms” is essential for their optimization and may even be necessary to obtain regulatory approval (25). To this end, eukaryotic expression systems have been developed that allow glycosylation to be controlled. Recently, Pichia pastoris-based strains with human glycosyltransferases have been established, allowing the expression of glycoforms with oligomannose-, hybrid-, and some complex-type glycans (26, 27) and even sialylated complex-type structures (28). However, mammalian expression remains the dominant technology in industrial settings, presumably because of its reliability for the expression of human secreted glycoproteins.Although the majority of lectin-resistant cell lines have been generated using CHO cells, no Golgi α-mannosidase II-deficient CHO cell line has been generated thus far (15). Furthermore, only one human lectin-resistant cell line, i.e. GnT I-deficient (Lec1) HEK 293S cells (29), has been produced. Hybrid-type glycosylation has been reported to accumulate in ricin-resistant baby hamster kidney cells (30, 31); however, these cells contain a reduced but detectable level of cell-surface complex-type glycans, consistent with an incomplete ablation of Golgi α-mannosidase II activity (31). Moreover, the hybrids from one of these lines contain a trimannosyl rather than pentamannosyl core and appear to be heavily influenced by GnT II deficiency, resulting in the formation of what are now commonly called monoantennary complex-type glycans (3234). We now describe the isolation of an HEK 293T cell line mutated at the MAN2A1 locus and deficient in Golgi α-mannosidase II activity via selection with ricin.  相似文献   

11.
Aberrant glycosylation of human tissue inhibitor of metalloproteinase-1 (TIMP-1) by N-acetylglucosaminyltransferase-V (GnT-V) was previously reported to be related to cancer progression. Here, we report on the antibodies recognizing the structural features initiated by an addition of N-linked β(1,6)-N-acetylglucosamine (GlcNAc) branch by GnT-V on TIMP-1. Two glycoforms of TIMP-1, TIMP1-L produced in Lec4 cells without GnT-V activity and TIMP1-B in GnT-V overexpressing transfectant cells, were purified from culture supernatant and used for generation of antibodies. TIMP1-L was injected in the left hind footpad of mice as decoy and TIMP1-B in the right hind footpad as immunogen. Phage-displayed scFv library was constructed from the B cells retrieved from the right popliteal lymph nodes and subjected to panning and screening. Phage ELISA of individual clones revealed the scFv clones with preferential binding activity to TIMP1-B, and they were converted into an scFv-Fc format for further characterization of binding specificity. Glycan binding assay of an antibody, 1-9F, revealed its differential specificity toward an extension of glycan structure initiated with β(1,6)-GlcNAc linkage and terminally decorated with a sialic acid. This study demonstrates feasibility of a new strategy combining decoy immunization with phage display for the efficient generation of antibodies tracking down structural features of different glycoforms.  相似文献   

12.
Novel chimeric lacdiNAc (GalNAc(β1-4)GlcNAc) synthase (c-LacdiNAcS) was generated by gene fusion of α-lactalbumin (α-LA) and β1,4-galactosyltransferase 1 (β1,4-GalT1). c-LacdiNAcS was expressed in Lec8 Chinese hamster ovary (Lec8 CHO) cells and exhibited N-acetylgalactosaminyltransferase (GalNAcT) activity in the absence of exogenous α-LA as well as other glycosyltransferase activities including lactose synthase (LacS), and β1,4-GalT. These glycosyltransferase activities of c-LacdiNAcS were compared to those activities induced in LacS system under the co-presence of bovine β1,4-GalT1 and α-LA, indicating that each domain of α-LA and β1,4-GalT1 on c-LacdiNAcS is not only folding correctly, but also interacting together. Furthermore, c-LacdiNAcS was found to be auto-lacdiNAcylated and can synthesize lacdiNAc structures on cellular glycoproteins, demonstrating that GalNAcT activity of c-LacdiNAcS is functional in Lec8 CHO cells.  相似文献   

13.
SLC35A3 is considered the main UDP-N-acetylglucosamine transporter (NGT) in mammals. Detailed analysis of NGT is restricted because mammalian mutant cells defective in this activity have not been isolated. Therefore, using the siRNA approach, we developed and characterized several NGT-deficient mammalian cell lines. CHO, CHO-Lec8, and HeLa cells deficient in NGT activity displayed a decrease in the amount of highly branched tri- and tetraantennary N-glycans, whereas monoantennary and diantennary ones remained unchanged or even were accumulated. Silencing the expression of NGT in Madin-Darby canine kidney II cells resulted in a dramatic decrease in the keratan sulfate content, whereas no changes in biosynthesis of heparan sulfate were observed. We also demonstrated for the first time close proximity between NGT and mannosyl (α-1,6-)-glycoprotein β-1,6-N-acetylglucosaminyltransferase (Mgat5) in the Golgi membrane. We conclude that NGT may be important for the biosynthesis of highly branched, multiantennary complex N-glycans and keratan sulfate. We hypothesize that NGT may specifically supply β-1,3-N-acetylglucosaminyl-transferase 7 (β3GnT7), Mgat5, and possibly mannosyl (α-1,3-)-glycoprotein β-1,4-N-acetylglucosaminyltransferase (Mgat4) with UDP-GlcNAc.  相似文献   

14.
The severe phenotypic effects of altered glycosylation in the congenital muscular dystrophies, including Walker-Warburg syndrome, muscle-eye-brain disease, Fukuyama congenital muscular dystrophy, and congenital muscular dystrophy 1D, are caused by mutations resulting in altered glycans linked to proteins through O-linked mannose. A glycosyltransferase that branches O-Man, N-acetylglucosaminyltransferase Vb (GnT-Vb), is highly expressed in neural tissues. To understand the expression and function of GnT-Vb, we studied its expression during neuromorphogenesis and generated GnT-Vb null mice. A paralog of GnT-Vb, N-acetylglucosaminyltransferase (GnT-V), is expressed in many tissues and brain, synthesizing N-linked, β1,6-branched glycans, but its ability to synthesize O-mannosyl-branched glycans is unknown; conversely, although GnT-Vb can synthesize N-linked glycans in vitro, its contribution to their synthesis in vivo is unknown. Our results showed that deleting both GnT-V and GnT-Vb results in the total loss of both N-linked and O-Man-linked β1,6-branched glycans. GnT-V null brains lacked N-linked, β1,6-glycans but had normal levels of O-Man β1,6-branched structures, showing that GnT-Vb could not compensate for the loss of GnT-V. By contrast, GnT-Vb null brains contained normal levels of N-linked β1,6-glycans but low levels of some O-Man β1,6-branched glycans. Therefore, GnT-V could partially compensate for GnT-Vb activity in vivo. We found no apparent change in α-dystroglycan binding of glycan-specific antibody IIH6C4 or binding to laminin in GnT-Vb null mice. These results demonstrate that GnT-V is involved in synthesizing branched O-mannosyl glycans in brain, but the function of these branched O-mannosyl structures is unresolved using mice that lack these glycosyltransferases.  相似文献   

15.
It is well documented that glycan synthesis is altered in some pathological processes, including cancer. The most frequently observed alterations during tumourigenesis are extensive expression of β1,6-branched complex type N-glycans, the presence of poly-N-acetyllactosamine structures, and high sialylation of cell surface glycoproteins. This study investigated two integrins, α3β1 and αvβ3, whose expression is closely related to cancer progression. Their oligosaccharide structures in two metastatic melanoma cell lines (WM9, WM239) were analysed with the use of matrix-assisted laser desorption ionisation mass spectrometry. Both examined integrins possessed heavily sialylated and fucosylated glycans, with β1,6-branches and short polylactosamine chains. In WM9 cells, α3β1 integrin was more variously glycosylated than αvβ3; in WM239 cells the situation was the reverse. Functional studies (wound healing and ELISA integrin binding assays) revealed that the N-oligosaccharide component of the tested integrins influenced melanoma cell migration on vitronectin and α3β1 integrin binding to laminin-5. Additionally, more variously glycosylated integrins exerted a stronger influence on these parameters. To the best of our knowledge, this is the first report concerning structural characterisation of αvβ3 integrin glycans in melanoma or in any cancer cells.  相似文献   

16.
Reichner  JS; Helgemo  SL; Hart  GW 《Glycobiology》1998,8(12):1173-1182
The ability of particular cell surface glycoproteins to recycle and become exposed to individual Golgi enzymes has been demonstrated. This study was designed to determine whether endocytic trafficking includes significant reentry into the overall oligosaccharide processing pathway. The Lec1 mutant of Chinese hamster ovary (CHO) cells lack N - acetylglucosaminyltransferase I (GlcNAc-TI) activity resulting in surface expression of incompletely processed Man5GlcNAc2 N -linked oligosaccharides. An oligosaccharide tracer was created by exoglycosylation of cell surface glycoproteins with purified porcine GlcNAc-TI and UDP-[3H]GlcNAc. Upon reculturing, all cell surface glycoproteins that acquired [3H]GlcNAc were acted upon by intracellular mannosidase II, the next enzyme in the Golgi processing pathway of complex N -linked oligosaccharides (t1/2= 3-4 h). That all radiolabeled cell surface glycoproteins were included in this endocytic pathway indicates a common intracellular compartment into which endocytosed cell surface glycoproteins return. Significantly, no evidence was found for continued oligosaccharide processing consistent with transit through the latter cisternae of the Golgi apparatus. These data indicate that, although recycling plasma membrane glycoproteins can be reexposed to individual Golgi-derived enzymes, significant reentry into the overall contiguous processing pathway is not evident.   相似文献   

17.
The transfer of N-acetylglucosamine (GlcNAc) to Ser or Thr in cytoplasmic and nuclear proteins is a well known post-translational modification that is catalyzed by the O-GlcNAc transferase OGT. A more recently identified O-GlcNAc transferase, EOGT, functions in the secretory pathway and transfers O-GlcNAc to proteins with epidermal growth factor-like (EGF) repeats. A number of antibodies that detect O-GlcNAc in cytosolic and nuclear extracts have been described previously. Here we compare seven of these antibodies (CTD110.6, 10D8, RL2, HGAC85, 18B10.C7(#3), 9D1.E4(#10), and 1F5.D6 (#14) for detection of the O-GlcNAc modification on extracellular domains of membrane or secreted glycoproteins that may also carry various N- and O-glycans. We found that CTD110.6 binds not only to O-GlcNAc on proteins but also to terminal β-GlcNAc on the complex N-glycans of Lec8 Chinese hamster ovary (CHO) cells that lack UDP-Gal transporter activity and express GlcNAc-terminating, complex N-glycans. We show that CTD110.6, #3, and #10 antibodies can be used to detect cell surface glycoproteins bearing O-GlcNAc. Cell surface glycoproteins recognized by CTD110.6 antibody included NOTCH1 that possesses many EGF repeats with a consensus site for EOGT. Knockdown of CHO Eogt reduced binding of CTD110.6 to Lec1 CHO cells, and expression of a human EOGT cDNA increased the O-GlcNAc signal on Lec1 cells and the extracellular domain of NOTCH1. Thus, with careful controls, antibodies CTD110.6 (IgM), #3 (IgG), and #10 (IgG) can be used to detect membrane and secreted proteins modified by O-GlcNAc on EGF repeats.  相似文献   

18.
Acquisition of metastatic potential is accompanied by changes in cell surface N-glycosylation. One of the best-studied changes is increased expression of N-acetylglucosaminyltransferase V enzyme (GnT-V) and its products, β1,6-branched N-linked oligosaccharides, observed in the tumorigenesis of many cancers. In this study we demonstrate that during the transition from the vertical growth phase (VGP) (WM793 cell line) to the metastatic stage (WM1205Lu line), β1,6 glycosylation of melanoma cell surface proteins increases as a consequence of elevated expression of the GnT-V-encoding Mgat-5 gene. Treatment with swainsonine led to reduced cell motility on fibronectin in both cell lines; the effect was stronger in metastatic cells, probably due to the higher content of GlcNAc β1,6-branched glycans on the main fibronectin receptors – integrins α5β1 and α3β1. Our results show that GlcNAc β1,6 N-glycosylation of cell surface receptors, which increases with the aggressiveness of melanoma cells, is an important factor influencing melanoma cell migration.  相似文献   

19.
NOTCH signaling induced by Delta1 (DLL1) and Jagged1 (JAG1) NOTCH ligands is modulated by the β3N-acetylglucosaminyl transferase Fringe. LFNG (Lunatic Fringe) and MFNG (Manic Fringe) transfer N-acetylglucosamine (GlcNAc) to O-fucose attached to EGF-like repeats of NOTCH receptors. In co-culture NOTCH signaling assays, LFNG generally enhances DLL1-induced, but inhibits JAG1-induced, NOTCH signaling. In mutant Chinese hamster ovary (CHO) cells that do not add galactose (Gal) to the GlcNAc transferred by Fringe, JAG1-induced NOTCH signaling is not inhibited by LFNG or MFNG. In mouse embryos lacking B4galt1, NOTCH signaling is subtly reduced during somitogenesis. Here we show that DLL1-induced NOTCH signaling in CHO cells was enhanced by LFNG, but this did not occur in either Lec8 or Lec20 CHO mutants lacking Gal on O-fucose glycans. Lec20 mutants corrected with a B4galt1 cDNA became responsive to LFNG. By contrast, MFNG promoted DLL1-induced NOTCH signaling better in the absence of Gal than in its presence. This effect was reversed in Lec8 cells corrected by expression of a UDP-Gal transporter cDNA. The MFNG effect was abolished by a DDD to DDA mutation that inactivates MFNG GlcNAc transferase activity. The binding of soluble NOTCH ligands and NOTCH1/EGF1-36 generally reflected changes in NOTCH signaling caused by LFNG and MFNG. Therefore, the presence of Gal on O-fucose glycans differentially affects DLL1-induced NOTCH signaling modulated by LFNG versus MFNG. Gal enhances the effect of LFNG but inhibits the effect of MFNG on DLL1-induced NOTCH signaling, with functional consequences for regulating the strength of NOTCH signaling.  相似文献   

20.
The effects of altering N-cadherin N-glycosylation on several cadherin-mediated cellular behaviors were investigated using small interfering RNA and site-directed mutagenesis. In HT1080 fibrosarcoma cells, small interfering RNA-directed knockdown of N-acetylglucosaminyltransferase V (GnT-V), a glycosyltransferase up-regulated by oncogene signaling, caused decreased expression of N-linked β(1,6)-branched glycans expressed on N-cadherin, resulting in enhanced N-cadherin-mediated cell-cell adhesion, but had no effect on N-cadherin expression on the cell surface. This effect on adhesion was accompanied by decreased cell migration and invasion, opposite of the effects observed when GnT-V was overexpressed in these cells (Guo, H. B., Lee, I., Kamar, M., and Pierce, M. (2003) J. Biol. Chem. 278, 52412–52424). A detailed study using site-directed mutagenesis demonstrated that three of the eight putative N-glycosylation sites in the N-cadherin sequence showed N-glycan expression. Moreover, all three of these sites, located in the extracellular domains EC2 and EC3, were shown by leucoagglutinating phytohemagglutinin binding to express at least some β(1,6)-branched glycans, products of GnT-V activity. Deletion of these sites had no effect on cadherin levels on the cell surface but led to increased stabilization of cell-cell contacts, cell-cell adhesion- mediated intracellular signaling, and reduced cell migration. We show for the first time that these deletions had little effect on formation of the N-cadherin-catenin complex but instead resulted in increased N-cadherin cis-dimerization. Branched N-glycan expression at three sites in the EC2 and -3 domains regulates N-cadherin-mediated cell-cell contact formation, outside-in signaling, and cell migration and is probably a significant contributor to the increase in the migratory/invasive phenotype of cancer cells that results when GnT-V activity is up-regulated by oncogene signaling.  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