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 46 毫秒
1.
Treatment of influenza virus and its purified hemagglutining with glycosidases from Diplococcus pneumoniae, which included beta-galactosidase, beta-N-acetylglucosminidase, and endoglycosidase D, released amino and neutral sugars from the virus and these as well as large oligosaccharides from the purified hemagglutinin. The released glucosamine-containing oligosaccharides were of one discrete size. Large oligosaccharides not removed by the glycosidases were found on the virus as well as the hemagglutinin. Some oligosaccharides on the virus were inaccessible to the enzymes, since they could be removed only from the purified hemagglutinin. Approximately 50% of the hemagglutinin carbohydrates could be removed without effect on hemagglutinating activity. Similarly, removal of 20 to 25% of the carbohydrates from intact virus particles did not alter infectivity.  相似文献   

2.
R Ohuchi  M Ohuchi  W Garten    H D Klenk 《Journal of virology》1997,71(5):3719-3725
The influenza A virus hemagglutinin (HA) has three conserved oligosaccharides located in the stem region at asparagine residues 12, 28, and 478. The biological role of these oligosaccharides has been investigated by mutational analysis of HA of fowl plague virus that was expressed from a simian virus 40 vector in the presence of ammonium chloride for protection from acid denaturation in the trans-Golgi network. Resistance to endoglycosidase H and cleavage of HA into the subunits HA1 and HA2 have been analyzed as markers for intracellular transport. Cell surface exposure has been determined by hemadsorption following neuraminidase treatment, by immunofluorescence staining, and by fluorescence-activated cell sorter analysis. When all three stem oligosaccharides were removed, transport was almost completely blocked. When two of the three stem oligosaccharides, particularly those at asparagine residues 12 and 28, were missing, HA was transported to the surface but showed extremely low fusion activity. With mutants lacking one stem oligosaccharide, fusion was reduced to a lesser extent. Removal of stem oligosaccharides resulted also in an increase in the pH optimum required for fusion. On the other hand, no reduction in fusion activity was observed when oligosaccharides in the head region of the HA spike were removed. These results indicate that the conserved oligosaccharides in the stem stabilize HA in the form susceptible to the conformational change necessary for fusion.  相似文献   

3.
The influence of cell transformation on the glycosylation of viral envelope glycoproteins was examined by high-resolution gel filtration and specific glycosidase digestions of 3H-sugar-labeled glycopeptides from nondefective and transformation-defective Prague C strains of Rous sarcoma virus replicated in fibroblasts from the same chicken embryo. The major difference in glycosylation attributable to the viral transformation of the host cells was an increase in this relative amount of larger acidic-type oligosaccharides containing additional "branch" sugars (NeuNAc-Gal-GlcNAc-) compared with the smaller acidic-type and neutral-type oligosaccharides. There was also a shift in size distribution of neutral-type oligosaccharides toward smaller oligomannosyl cores in the transforming versus nontransforming virus glycopeptides. These alterations were consistent with a transformation-dependent increase in the extent of intracellular processing of a common precursor structure for the asparagine-linked oligosaccharides of Rous sarcoma virus.  相似文献   

4.
Because of the extensive oligosaccharide heterogeneity of the membrane glycoprotein (G) from the Hazelhurst strain of vesicular stomatitis virus, this virus has been used as a specific intracellular probe of altered protein glycosylation in Rous sarcoma virus-transformed versus normal baby hamster kidney cells. Over 70% of G protein from virus released from the transformed cells had acidic-type oligosaccharides at both glycosylation sites, compared to less than 50% from the corresponding normal host cells. The remaining G protein contained an acidic-type oligosaccharide at one site and an endo-beta-N-acetylglucosaminidase H-sensitive oligosaccharide at the other. The major endoglycosidase-sensitive species were sialylated hybrid-type (NeuNAc-Gal-GlcNAc-Man5GlcNAc2-Asn) from the transformed and neutral-type (Man5-6GlcNAc2-Asn) from the normal host cells. The degree of branching of the acidic-type oligosaccharides was not increased in the transformed cells (approx. 80% biantennary for viral G protein from both cell types). At a reduced growth temperature (24 versus 37 degrees C), the G protein oligosaccharides were more extensively processed in both cell types (approximately 85-95% of G protein contained acidic-type structures at both sites), even though the level of viral protein synthesis and virus release was decreased. Essentially all of the minor, endoglycosidase-sensitive oligosaccharides on mature viral G protein were sialic acid-containing hybrid-type structures. At 24 degrees C the branching of the acidic-type oligosaccharides was increased in the virus released from the transformed cells versus normal cells.  相似文献   

5.
D H Madoff  J Lenard 《Cell》1982,28(4):821-829
The intracellular transport and certain posttranslational modifications of the large glycoprotein (G1) of LaCrosse virus (LAC) in BHK cells have been studied. G1 from released LAC virus was characterized by complex oligosaccharides (endo H-resistant) and covalently attached fatty acid. Only a small fraction of total cellular G1 was present on the baby hamster kidney cell surface. Cell-surface G1 contained complex oligosaccharides, while total G1 in infected cells contained largely unprocessed (endo H-sensitive) oligosaccharides. In addition, cell G1 contained significantly less fatty acid than virion-associated G1. Pulse-chase experiments showed that the oligosaccharides of G1 were processed to the complex from much more slowly than the oligosaccharides of the vesicular stomatitis virus (VSV) glycoprotein (G). In addition, transit of LAC G1 to the cell surface and into extracellular virions was two to three fold slower than the transit of VSV G. Thus LAC G1 accumulates intracellularly and is only slowly processed by intracellular processing enzymes. Treatment with monensin caused accumulation in the cell of a form of G1 with partial sensitivity toward endo H, suggesting that monensin may act to inhibit the glycosylation process directly.  相似文献   

6.
Intact Sindbis virus and Triton-solubilized viral glycoprotein were treated with alpha-mannosidase and with a preparation of mixed glycosidases from Diplococcus pneumoniae to probe the accesibility of carbohydrate units on the viral surface. The products of glycosidase attack on Triton-solubilized virus showed that mose carbohydrate units of the glycoproteins are good substrates for these enzymes. The relative resistance of most of the viral oligosaccharides in intact virus particles showed that much of the carbohydrate is not accessible to glycosidases, probably because it is not exposed at the viral surface. The only completely accessible carbohydrate units on Sindbis glycoproteins were the type A oligosaccharides of E2. This differential accessibility of Sindbis oligosaccharides is discussed in relation to the organization of the viral surface.  相似文献   

7.
The oligosaccharides of the membrane glycoproteins of Sindbis virus, vesicular stomatitis virus, and Rous sarcoma virus were compared on the basis of apparent size and sugar composition. It appears that each virus acquires a different set of oligosaccharides during growth in a single type of cell.  相似文献   

8.
D L Huso  O Narayan    G W Hart 《Journal of virology》1988,62(6):1974-1980
The lentivirus caprine arthritis-encephalitis virus (CAEV) is a pathogen of goats. It is transmitted in milk and causes a persistent infection in goats, which often fail to produce neutralizing antibodies to the virus. Native CAEV particles are remarkably resistant to digestion with proteinase K and are neutralized extremely slowly by immune sera. Our studies showed that the virus particles are heavily sialylated. Studies with highly specific sialyltransferase enzymes identified penultimate carbohydrate linkages typical of O- and N-linked oligosaccharides on the virus and suggested that the virus may be more heavily sialylated on O-linked than on N-linked oligosaccharides. Removal of sialic acids from the virus by neuraminidase treatment did not reduce infectivity of the particles. However, desialylation rendered the virus more susceptible to proteolysis by proteinase K. Desialylation also enhanced the kinetics of neutralization of the virus by goat antibodies. These results suggest that the carbohydrates on the viral surface are important both in protecting viral proteins from digestion by proteases and in protecting the virus from rapid neutralization by antibodies.  相似文献   

9.
We have previously reported that the binding properties of the hemagglutinin (HA) of the WSN-F strain of influenza A are affected by the cells in which the virus is grown (Crecelius, D. M., Deom, C. M., and Schulze, I.T. (1984) Virology 139, 164-177); at 37 degrees C chick embryo fibroblast-grown F virus has a greater affinity for host cells than does the same virus grown in Madin-Darby bovine kidney (MDBK) cells. In an attempt to explain this host-determined property, we have characterized the carbohydrate put onto the viral HA by these two cells. Experiments using tunicamycin indicate that the HA made by MDBK cells contains about 4000 daltons of carbohydrate in excess of that on the HA from chick embryo fibroblast. Serial lectin affinity chromatography of the asparagine-linked oligosaccharides on the HA subunits, HA1 and HA2, detected a number of host-dependent differences in the complex oligosaccharides. Both HA1 and HA2 from MDBK cells contained more highly branched (i.e. tri- and tetraantennary) complex oligosaccharides than did the subunits from chick embryo fibroblasts. In addition, the HA subunits from the two sources differed in the amount of galactose-containing "bisected" complex oligosaccharides and in the presence of certain fucosylated triantennary oligosaccharides. Profiles of the asparagine-linked oligosaccharides from the host cells did not show these differences, indicating that the HA subunit profiles were not necessarily representative of the structures found on the cellular glycoproteins. The data support the conclusion that bulky oligosaccharides on the MDBK-HA subunits of WSN-F reduce the affinity of the virus for cellular receptors.  相似文献   

10.
The asparagine-linked oligosaccharides of the G protein of the Hazelhurst subtype of the New Jersey serotype of vesicular stomatitis virus (VSV) have been compared with the oligosaccharides from the G protein of the well-characterized Indiana serotype of VSV, with baby hamster kidney cells in monolayer culture as the host for both viruses. [3H]Glucosamine- and [3H]mannose-labeled glycopeptides from the G protein of purified virus were analyzed by the combined techniques of endo-beta-N-acetylglucosaminidase H (ENDO-H) digestion, concanavalin A and lentil lectin affinity chromatography, and Bio-Gel P-4 chromatography. Although almost all of the Indiana G protein oligosaccharides were acidic-type structures, as expected from previous studies; the Hazelhurst G protein contained a mixture of acidic-type, hybrid-type containing sialic acid, and neutral-type (predominantly Man5-6GlcNAc2-Asn) structures. The vast majority of acidic-type oligosaccharides from both the Hazelhurst and Indiana G proteins were diantennary structures, with less than half containing fucose linked to the innermost N-acetylglucosamine. Additional analysis of the Hazelhurst G protein by ENDO-H digestion and gel electrophoresis suggested that some of the mature G polypeptides contained acidic-type structures at both glycosylation sites, whereas the remainder contained an ENDO-H-resistant, acidic-type structure at one site and an ENDO-H-sensitive, hybrid- or neutral-type structure at the other site.  相似文献   

11.
By pulse-chase labeling with [35S]methionine and long-term labeling with 3H-sugars, the E1 glycoprotein of coronavirus MHV-A59 has been shown to acquire O-linked oligosaccharides in a two-step process. About 10 min after synthesis of the E1 protein, N-acetyl-galactosamine was added. This was followed approximately 10 min later by the addition of both galactose and sialic acid to give the mature oligosaccharides. This sequence of additions was confirmed by analyzing the 3H-labeled oligosaccharides bound to each of the E1 forms using gel filtration on P4 columns. The intracellular location of the first step was determined by exploiting the temperature sensitivity of virus release. The virus normally buds first into a smooth membrane compartment lying between the rough endoplasmic reticulum and the cis side of the Golgi stack (Tooze et al., 1984). At 31 degrees C the virus is assembled but does not appear to enter the Golgi stacks. The addition of N-acetyl-galactosamine is unaffected although the addition of galactose and sialic acid is inhibited. These results strongly suggest that addition of N-acetyl-galactosamine occurs in this budding compartment, the morphology of which is similar to that of transitional elements and vesicles.  相似文献   

12.
M Ohuchi  R Ohuchi  A Feldmann    H D Klenk 《Journal of virology》1997,71(11):8377-8384
The hemagglutinin (HA) of the fowl plague virus (FPV) strain of influenza A virus has two N-linked oligosaccharides attached to Asn123 and Asn149 in the vicinity of the receptor binding site. The effect of these carbohydrate side chains on the binding of HA to neuraminic acid-containing receptors has been analyzed. When the oligosaccharides were deleted by site-specific mutagenesis, HA expressed from a simian virus 40 vector showed enhanced hemadsorbing activity. Binding was so strong under these conditions that erythrocytes were no longer released by viral neuraminidase and that release was significantly reduced when neuraminidase from Vibrio cholerae was used. Similarly, when these oligosaccharides were removed selectively from purified viruses by N-glycosidase F, such virions were unable to elute from receptors, although they retained neuraminidase activity. Thus, release of FPV from cell receptors depends on the presence of the HA glycans at Asn123 and Asn149. On the other hand, receptor binding was abolished when these oligosaccharides were sialylated after expression in the absence of neuraminidase (M. Ohuchi, A. Feldmann, R. Ohuchi, and H.-D. Klenk, Virology 212:77-83, 1995). These observations indicate that the receptor affinity of FPV HA is controlled by oligosaccharides adjacent to the receptor binding site.  相似文献   

13.
The main surface glycoprotein, hemagglutinin (HA), was obtained by treatment of influenza virus B/Leningrad/179/86 with bromelain. Amino acid and monosaccharide compositions of HA and neuraminidase (NA, earlier isolated from the same virus) were determined, thus showing HA and NA to contain 8-10 and 2 carbohydrate chains, respectively. The carbohydrate fragments were cleaved off by the alkaline LiBH4 treatment, the oligosaccharides released were reduced with NaB3H4 and fractionated by two-step HPLC on Ultrasphere-C18 and Zorbax-NH2 columns. Some higher mannose and complex oligosaccharides were identified in both cases by comparison with nonlabelled oligosaccharides of the known structure. The data obtained show that surface glycoproteins of influenza virus A and B are rather similar with regard to structure and heterogeneity of their carbohydrate chains.  相似文献   

14.
Two cellular proteins, membrane cofactor protein (MCP) and moesin, were reported recently to be functionally associated with the initiation of a measles virus infection. We have analyzed the interaction of measles virus with cell surface proteins, using an overlay binding assay with cellular proteins immobilized on nitrocellulose. Among surface-biotinylated proteins from a human rectal tumor cell line (HRT), measles virus was able to bind only to a 67-kDa protein that was identified as MCP. The virus recognized different isoforms of MCP expressed from human (HRT and HeLa) and simian (Vero) cell lines. The binding of measles virus to MCP was abolished after cleavage of the disulfide bonds by reducing agents as well as after enzymatic release of N-linked oligosaccharides. By contrast, removal of sialic acid or O-linked oligosaccharides did not affect the recognition of MCP measles virus. These data indicate that the receptor determinant of MCP is dependent on a conformation of the protein that is maintained by disulfide bonds and N-glycans present in the complement binding domains. Our results are consistent with a role of MCP as primary attachment site for measles virus in the initial stage of an infection. The functional relationship between MCP and moesin in a measles virus infection is discussed.  相似文献   

15.
In comparisons of [3H]mannose-labelled glycopeptides from chick-embryo fibroblasts infected and transformed with non-defective Prague C Rous-sarcoma virus and from untransformed fibroblasts infected with a transformation-defective derivative of Prague C Rous-sarcoma virus, we have detected transformation-dependent alterations in both the acidic-type and the neutral-type asparagine-linked oligosaccharides of cellular glycoproteins. Pronase-digested glycopeptides were analysed by the combined techniques of gel filtration, exo- and endo-glycosidase digestion and concanavalin A-agarose affinity chromatography. The transformed cell glycoproteins contained more sialic acid and were enriched for more highly branched (versus biantennary) acidic-type structures compared with the untransformed cell glycoproteins, similarly to previously reported transformation-dependent alterations. In addition, the glycopeptides from the virus-transformed cells contained several neutral-type structures that were apparently absent from the untransformed cells: small neutral-type oligosaccharides (Man3GlcNAc2) that were sensitive to endo-beta-N-acetylglucosaminidase D but resistant to endo-beta-N-acetylglucosaminidase H, and oligosaccharides with the property of 'truncated' precursor oligosaccharides (endoglycosidase-resistant, alpha-mannosidase-sensitive). Endoglycosidase-released oligosaccharides with the properties of hybrid-type structures were derived from the glycoproteins of both transformed and untransformed cells.  相似文献   

16.
The structures of the asparagine-linked oligosaccharides of several variant forms of the vesicular stomatitis virus glycoprotein transiently expressed from cloned cDNAs have been determined. Glycopeptides isolated from forms of the G protein that reach the cell surface or that are secreted into the medium are virtually identical; they contain complex-type oligosaccharides whose nonreducing ends terminate in galactose and sialic acid residues. In contrast, forms of the G protein that remain intracellular possess oligosaccharides at intermediate stages in the processing pathway. One deletion mutant, delta 1473, codes for a protein that remains in the rough endoplasmic reticulum (Rose, J. K., and J. E. Bergmann, 1982, Cell, 30:753-762) and contains only high mannose-type oligosaccharides. Another mutant, delta 1554, codes for a glycoprotein that contains oligosaccharides of primarily two classes. One class is of the high mannose type and is similar to those found on the protein coded for by delta 1473. However, the major class contains biantennary and more highly branched complex-type oligosaccharides that terminate in N-acetylglucosamine rather than galactose or sialic acid residues. These data suggest that the protein coded for by delta 1554 migrates to the Golgi apparatus, but does not enter the more distal compartment(s) of the organelle which contains galactosyl- and sialyltransferases.  相似文献   

17.
The role of glucosylated oligosaccharides in the biogenesis of the glycoprotein (G protein) of vesicular stomatitis virus was studied in PhaR2.7, a mouse lymphoma cell line deficient in glucosidase II activity. As expected, the great majority of cell-associated G protein remained glucosylated in PhaR2.7, and the G protein was rapidly deglucosylated in BW5147, the parental cell line. Despite these differences in glucosylation, the rates of G protein trimerization and transport to the cell surface were as rapid and efficient in the PhaR2.7 mutant as in BW5147. Surprisingly, greater than 73% of the oligosaccharides on G proteins recovered from released virions were complex-type units. The efficient processing of the G protein oligosaccharides coincided with the efficient removal of glucose residues from its oligosaccharides. After treatment with deoxynojirimycin, an inhibitor of endoplasmic reticulum (ER) glucosidases I and II, the total percentage of G protein-associated high mannose-type oligosaccharides increased more in the parental cells than in the mutant cells. Furthermore, when the G protein was retained in the ER of PhaR2.7 cells by depletion of the cellular ATP pools with carbonyl cyanide m-chlorophenylhydrazone, its oligosaccharides remained glucosylated. Under identical conditions, BW5147 cells removed the glucose residues from > 90% of the retained G protein's oligosaccharides. Thus, PhaR2.7 cells efficiently remove glucose residues from high mannose-type oligosaccharides of selected proteins using a deoxynojirimycin-insensitive enzyme located in a post-ER compartment. The existence of a second mechanism for the deglucosylation of N-linked oligosaccharides provides evidence for the important role of glucose removal in glycoprotein maturation.  相似文献   

18.
The O-linked oligosaccharides on mature forms of herpes simplex virus type 1 (HSV1) glycoproteins were characterized, and were found to account largely for the lower electrophoretic mobilities of these forms relative to the mobilities of immature forms. Other posttranslational modifications of HSV1 glycoproteins (designated gB, gC, gD and gE) were related temporally to the discrete shifts in electrophoretic mobilities that signal acquisition of the O-linked oligosaccharides. Fatty acid acylation (principally of gE) could be detected just prior to the shifts, whereas conversion of high-mannosetype N-linked oligosaccharides to the complex type occurred coincident with the shifts. The addition of O-linked oligosaccharides did not occur in cells treated with the ionophore monensin or in a ricinresistant cell line defective in the processing of N-linked oligosaccharides. We conclude that extension of O-linked oligosaccharide chains on HSV1 glycoproteins, and probably also attachment of the first O-linked sugars, occurs as a late posttranslational modification in the Golgi apparatus.  相似文献   

19.
We have examined and compared the host-cell-dependent glycosylation of the G glycoprotein of vesicular-stomatitis virus (Hazelhurst strain) and the E1 and E2 glycoproteins of Sindbis virus replicated by baby-hamster kidney, chicken-embryo fibroblast and mouse L929 monolayer cell cultures. The results of endo-beta-N-acetylglucosaminidase H digestion of viral proteins labelled with [3H]mannose or leucine and Pronase-digested glycopeptides labelled with [3H]mannose indicated that both the G protein and the E1 protein contained a similar mixture of endoglycosidase-resistant oligosaccharides of the complex acidic type and less extensively processed endoglycosidase-sensitive oligosaccharides of the neutral or hybrid type, with a relatively greater content of the endoglycosidase-sensitive oligosaccharides for virus replicated in the chicken as against hamster or mouse cells. A large fraction of the G protein and the majority of the E1 proteins from the mammalian host cells contained acidic-type oligosaccharides at both glycosylation sites, whereas most of the G and E1 glycoproteins from the avian host cells and essentially all of the E2 protein from all three host-cell types contained an acidic-type oligosaccharide at one site and neutral- or hybrid-type oligosaccharide at the other site. The relative increase in neutral- and hybrid-type oligosaccharides with five-mannose core structures observed for the G and E1 proteins of virus released from the avian host cells suggested that two specific steps in oligosaccharide processing (mediated by alpha-mannoside II and N-acetylglucosaminyltransferase I) were less efficient at one of the glycosylation sites of the vesicular-stomatitis-virus G protein and Sindbis-virus E1 protein in the avian as against mammalian host cells.  相似文献   

20.
The recent finding that the E1 glycoproteins of murine coronaviruses contain only O-linked oligosaccharides suggested that this unusual modification might be a distinguishing feature of coronaviruses and might play an essential role in the life cycle of this family of viruses. To examine these possibilities, we analyzed the oligosaccharide moieties of the membrane proteins of the avian coronavirus infectious bronchitis virus. In addition, we determined the effect of inhibiting the glycosylation of these proteins on viral maturation and infectivity. Infectious bronchitis virus virions contain nine proteins. Four of these proteins, GP36, GP31, GP28, and P23, are closely related structurally and appear to be homologous to the E1 proteins of murine coronaviruses. We found that the oligosaccharides of GP31 and GP28 could be removed with endoglycosidase H and that neither of these glycoproteins was detectable in tunicamycin-treated cells. These two results indicated that GP31 and GP28 contain N-linked oligosaccharides. Therefore, O-linked oligosaccharides are not a universal feature of the small coronavirus membrane glycoproteins. Tunicamycin inhibited glycosylation of all of the viral glycoproteins but did not inhibit production of virions by infectious bronchitis virus-infected cells. The virions released by these cells contained only the three non-glycosylated viral proteins P51, P23, and P14. These particles were not infectious. Therefore, it appears that glycosylated infectious bronchitis virus polypeptides are not required for particle formation. However, the viral glycoproteins are apparently indispensible for viral infectivity.  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