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 562 毫秒
1.
Here, we propose a novel method for the discrimination of α2,3- and α2,6-sialylation on glycopeptides. To stabilize the sialic acids, the carboxyl moiety on the sialic acid as well as the C-terminus and side chain of the peptide backbone were derivatized using 1-pyrenyldiazomethane (PDAM). The derivatization can be performed on the target plate for matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS), thereby avoiding complicated and time-consuming purification steps. After the on-plate PDAM derivatization, samples were subjected to negative-ion MALDI-MS using 3AQ-CHCA as a matrix. Deprotonated ions of the PDAM-derivatized form were detected as the predominant species without loss of sialic acid. The negative-ion collision-induced dissociation (CID) of PDAM-derivatized isomeric sialylglycopeptides, derived from hen egg yolk, showed characteristic spectral patterns. These data made it possible to discriminate α2,3- and α2,6-sialylation. In addition, sialyl isomers of a glycan with an asparagine could be discriminated based on their CID spectra. In brief, the negative-ion CID of PDAM-derivatized glycopeptides with α2,6-sialylation gave an abundant (0,2)A-type product ion, while that with α2,3-sialylation furnished a series of (2,4)A/Y-type product ions with loss of sialic acids. The unique fragmentation behavior appears to be derived from the difference of pyrene binding positions after ionization, depending on the type of sialylation. Thus, we show that on-plate PDAM derivatization followed by negative-ion MALDI-MS(2) is a simple and robust method for the discrimination of α2,3- and α2,6-sialylation on glycopeptides.  相似文献   

2.
Herpes simplex virus type 1 (HSV-1) envelope proteins are posttranslationally modified by the addition of sialic acids to the termini of the glycan side chains. Although gC, gD, and gH are sialylated, it is not known whether sialic acids on these envelope proteins are functionally important. Digestion of sucrose gradient purified virions for 4 h with neuraminidases that remove both alpha2,3 and alpha2,6 linked sialic acids reduced titers by 1,000-fold. Digestion with a alpha2,3-specific neuraminidase had no effect, suggesting that alpha2,6-linked sialic acids are required for infection. Lectins specific for either alpha2,3 or alpha2,6 linkages blocked attachment and infection to the same extent. In addition, the mobility of gH, gB, and gD in sodium dodecyl sulfate-polyacrylamide gel electrophoresis gels was altered by digestion with either alpha2,3 specific neuraminidase or nonspecific neuraminidases, indicating the presence of both linkages on these proteins. The infectivity of a gC-1-null virus, DeltagC2-3, was reduced to the same extent as wild-type virus after neuraminidase digestion, and attachment was not altered. Neuraminidase digestion of virions resulted in reduced VP16 translocation to the nucleus, suggesting that the block occurred between attachment and entry. These results show for the first time that sialic acids on HSV-1 virions play an important role in infection and suggest that targeting virion sialic acids may be a valid antiviral drug development strategy.  相似文献   

3.
We have previously shown that costimulation of endothelial cells with IL-1 + IL-4 markedly inhibits VCAM-1-dependent adhesion under flow conditions. We hypothesized that sialic acids on the costimulated cell surfaces may contribute to the inhibition. Northern blot analyses showed that Gal beta 1-4GlcNAc alpha 2, 6-sialyltransferase (ST6N) mRNA was up-regulated in cultured HUVEC by IL-1 or IL-4 alone, but that the expression was enhanced by costimulation, whereas the level of Gal beta 1-4GlcNAc/Gal beta 1-3GalNAc alpha2,3-sialyltransferase (ST3ON) mRNA was unchanged. Removing both alpha 2,6- and alpha 2,3-linked sialic acids from IL-1 + IL-4-costimulated HUVEC by sialidase significantly increased VCAM-1-dependent adhesion, whereas removing alpha 2,3-linked sialic acid alone had no effect; adenovirus-mediated overexpression of ST6N with costimulation almost abolished the adhesion, which was reversible by sialidase. The same treatments of IL-1-stimulated HUVEC had no effect. Lectin blotting showed that VCAM-1 is decorated with alpha 2,6- but not alpha 2,3-linked sialic acids. However, overexpression of alpha 2,6-sialyltransferase did not increase alpha 2,6-linked sialic acid on VCAM-1 but did increase alpha 2,6-linked sialic acids on other proteins that remain to be identified. These results suggest that alpha 2,6-linked sialic acids on a molecule(s) inducible by costimulation with IL-1 + IL-4 but not IL-1 alone down-regulates VCAM-1-dependent adhesion under flow conditions.  相似文献   

4.
Receptor recognition and binding is the first step in the viral cycle. It has been established that Newcastle Disease Virus (NDV) interacts with sialylated molecules such as gangliosides and glycoproteins at the cell surface. Nevertheless, the specific receptor(s) that mediate virus entry are not well known. We have analysed the role of the sialic acid linkage in the early steps of the viral infection cycle. Pretreatment of ELL-0 cells with both α2,3 and α2,6 specific sialidases led to the inhibition of NDV binding, fusion and infectivity, which were restored after α2,3(N)- and α2,6(N)-sialyltransferase incubation. Moreover, α2,6(N)-sialyltransferases also restored NDV activities in α2-6-linked sialic acid deficient cells. Competition with α2-6 sialic acid-binding lectins led to a reduction in the three NDV activities (binding, fusion and infectivity) suggesting a role for α2-6- linked sialic acid in NDV entry. We conclude that both α2-3- and α2-6- linked sialic acid containing glycoconjugates may be used for NDV infection. NDV was able to efficiently bind, fuse and infect the ganglioside-deficient cell line GM95 to a similar extent to that of its parental MEB4, suggesting that gangliosides are not essential for NDV binding, fusion and infectivity. Nevertheless, the fact that the interaction of NDV with cells deficient in N-glycoprotein expression such as Lec1 was less efficient prompted us to conclude that NDV requires N-linked glycoproteins for efficient attachment and entry into the host cell.  相似文献   

5.
The surface of vascular endothelium bears a glycocalyx comprised, in part, of a complex mixture of oligosaccharide chains attached to cell-surface proteins and membrane lipids. Importantly, understanding of the structure and function of the endothelial glycocalyx is poorly understood. Preliminary studies have demonstrated structural differences in the glycocalyx of pulmonary artery endothelial cells compared with pulmonary microvascular endothelial cells. Herein we begin to probe in more detail structural and functional attributes of endothelial cell-surface carbohydrates. In this study we focus on the expression and function of sialic acids in pulmonary endothelium. We observed that, although pulmonary microvascular endothelial cells express similar amounts of total sialic acids as pulmonary artery endothelial cells, the nature of the sialic acid linkages differs between the two cell types such that pulmonary artery endothelial cells express both α(2,3)- and α(2,6)-linked sialic acids on the surface (i.e., surficially), whereas microvascular endothelial cells principally express α(2,3)-linked sialic acids. To determine whether sialic acids play a role in endothelial barrier function, cells were treated with neuraminidases to hydrolyze sialic acid moieties. Disruption of cell-cell and cell-matrix adhesions was observed following neuraminidase treatment, suggesting that terminal sialic acids promote endothelial barrier integrity. When we measured transendothelial resistance, differential responses of pulmonary artery and microvascular endothelial cells to neuraminidase from Clostridium perfringens suggest that the molecular architecture of the sialic acid glycomes differs between these two cell types. Collectively our observations reveal critical structural and functional differences of terminally linked sialic acids on the pulmonary endothelium.  相似文献   

6.
Three O-acylated, unsaturated sialic acids, N-acetyl-9-O-acetyl-, N-acetyl-9-O-lactoyl-, and 2-deoxy-N-glycoloyl-9-O-lactoyl-2,3-didehydroneuraminic acid (5-acetamido-9-O-acetyl-, 5-acetamido-9-O-lactoyl-, and 2,6-anhydro-3,5-dideoxy-5-glycoloylamido-9-O-lactoyl-D-glycero-D-g alacto-non-2- enonic acid) were isolated from urine or submandibular glands of rat, pig, and cow. Mass spectrometric evidence for the existence of 2,3-unsaturated 9-O-acetyl-N-glycoloylneuraminic acid in porcine urine was also obtained. The sialic acids were purified by dialysis, gel- and ion-exchange chromatography, and preparative thin-layer chromatography. They were analyzed by thin-layer chromatography, high-pressure liquid chromatography, and capillary gas-liquid chromatography-mass spectrometry. For comparison, O-acetylated unsaturated sialic acids were synthesized.  相似文献   

7.
We used lectin cytochemistry and confocal microscopy to examine the distribution of sialic acid in epithelial cells. Maackia amurensis lectin and Sambuccus nigra agglutinin were used to detect alpha2,3 and alpha2,6 sialic acid, respectively. In Caco-2, HT-29 5M12, and MCF-7 cells, which express sialic acid mainly in one type of linkage, the majority of the signal was observed in the apical membrane. In cells that bound both lectins, alpha2,3 sialic acid was distributed apically, whereas alpha2,6 sialic acid showed a broader distribution. In IMIM-PC-1 cultures, alpha2,3 sialic acid was detected mainly in the apical membrane, whereas alpha2,6 sialic acid was more abundant in the basolateral domain of polarized cells. In these cells, treatment with GalNAc-O-benzyl led to reduced alpha2,3 levels and to an increase and redistribution of alpha2,6 to the apical domain. Similarly, sialic acid was predominantly expressed apically in all epithelial tissues examined. In conclusion, (a) sialic acid is mainly distributed to the apical membrane of epithelial cells, (b) there is a hierarchy in the distribution of sialic acids in polarized epithelial cells, i.e., alpha2,3 is preferred to alpha2,6 in the apical membrane, and (c) IMIM-PC-1 cells are a good model in which to study the regulation of the levels and distribution of sialic acids.  相似文献   

8.
Chinese hamster ovary (CHO) cells typically produce glycoproteins with N-glycans terminating in α-2,3 sialylation. Human cells produce glycoproteins that include α-2,3 and α-2,6 sialic acids. To examine the impact of altering protein sialylation on pharmacokinetic properties, recombinant human butyrylcholinesterase (BChE) was produced in CHO cells by knocking out the α-2,3 sialyltransferase genes followed by overexpression of the α-2,6 sialyltransferase (26BChE) enzyme. The N-glycan composition of 26BChE was compared to BChE with α-2,3 sialylation (23BChE) derived from wild-type CHO cells. Both 23BChE and 26BChE exhibited comparable antennarity distributions with bi-antennary di-sialylated glycans representing the most abundant glycoform. CD-1 mice were intravenously injected with the 23BChE or 26BChE, and residual BChE activities from blood collected at various time points for pharmacokinetic analyses. Although 23BChE contained a slightly lower initial sialylation level compared to 26BChE, the molecule exhibited higher residual activity between 5 and 24 hr postinjection. Pharmacokinetic analyses indicated that 23BChE exhibited an increase in area under the curve and a lower volume of distribution at steady state than that of 26BChE. These findings suggest that the type of sialylation linkage may play a significant role in the pharmacokinetic behavior of a biotherapeutic when tested in in vivo animal models.  相似文献   

9.
《MABS-AUSTIN》2013,5(8):1381-1390
ABSTRACT

Human IgG antibodies containing terminal alpha 2,6-linked sialic acid on their Fc N-glycans have been shown to reduce antibody-dependent cell-mediated cytotoxicity and possess anti-inflammatory properties. Although terminal sialylation on complex N-glycans can happen via either an alpha 2,3-linkage or an alpha 2,6-linkage, sialic acids on human serum IgG Fc are almost exclusively alpha 2,6-linked. Recombinant IgGs expressed in Chinese hamster ovary (CHO) cells, however, have sialic acids through alpha 2,3-linkages because of the lack of the alpha 2,6-sialyltransferase gene. The impact of different sialylation linkages to the structure of IgG has not been determined. In this work, we investigated the impact of different types of sialylation to the conformational stability of IgG through hydrogen/deuterium exchange (HDX) and limited proteolysis experiments. When human-derived and CHO-expressed IgG1 were analyzed by HDX, sialic acid-containing glycans were found to destabilize the CH2 domain in CHO-expressed IgG, but not human-derived IgG. When structural isomers of sialylated glycans were chromatographically resolved and identified in the limited proteolysis experiment, we found that only alpha 2,3-linked sialic acid on the 6-arm (the major sialylated glycans in CHO-expressed IgG1) destabilizes the CH2 domain, presumably because of the steric effect that decreases the glycan-CH2 domain interaction. The alpha 2,6-linked sialic acid on the 3-arm (the major sialylated glycan in human-derived IgG), and the alpha 2,3-linked sialic acid on the 3-arm, do not have this destabilizing effect.  相似文献   

10.
Transfer of terminal alpha 2,6-linked sialic acids to N-glycans is catalyzed by beta-galactoside alpha 2,6-sialyltransferase (ST6Gal I). Expression of ST6Gal I and its products is reportedly increased in colon cancers. To investigate directly the functional effects of ST6Gal I expression, human colon cancer (HT29) cells were transfected with specific antisense DNA. ST6Gal I mRNA and protein were virtually undetectable in six strains of transfected HT29 cells. ST6Gal activity was reduced to 14% of control (P<0.005) in transfected cells. Expression of terminal alpha 2,6- and alpha 2,3-linked sialic acids, and unmasked N-acetyllactosamine oligosaccharides, respectively, was assessed using flow cytometry and fluoresceinated Sambucus nigra, Maackia amurensis and Erythrina cristagalli lectins. Results indicated a major reduction in expression of alpha 2,6-linked sialic acids and counterbalancing increase in unmasked N-acetyllactosamines in antisense DNA-transfected cells, without altered expression of alpha 2,3-linked sialic acids or ganglioside profiles. The ability of transfected cells to form colonies in soft agar and to invade extracellular matrix material (Matrigel), respectively, in vitro was reduced by approx. 98% (P<0.0001) and more than 3-fold (P<0.005) compared to parental HT29 cells. These results indicate that N-glycans bearing terminal alpha 2,6-linked sialic acids may enhance the invasive potential of colon cancer cells.  相似文献   

11.
Chinese hamster ovary (CHO) cells are widely employed to produce glycosylated recombinant proteins. Our group as well as others have demonstrated that the sialylation defect of CHO cells can be corrected by transfecting the α2,6-sialyltransferase (α2,6-ST) cDNA. Glycoproteins produced by such CHO cells display both α2,6- and α2,3-linked terminal sialic acid residues, similar to human glycoproteins. Here, we have established a CHO cell line stably expressing α2,6-ST, providing a universal host for further transfections of human genes. Several relevant parameters of the universal host cell line were studied, demonstrating that the α2,6-ST transgene was stably integrated into the CHO cell genome, that transgene expression was stable in the absence of selective pressure, that the recombinant sialyltransferase was correctly localized in the Golgi and, finally, that the bioreactor growth parameters of the universal host were comparable to those of the parental cell line. A second step consisted in the stable transfection into the universal host of cDNAs for human glycoproteins of therapeutic interest, i.e. interferon-γ and the tissue inhibitor of metalloproteinases-1. Interferon-γ purified from the universal host carried 40.4% α2,6- and 59.6% α2,3-sialic acid residues and showed improved pharmacokinetics in clearance studies when compared to interferon-γ produced by normal CHO cells.  相似文献   

12.
Recombinant adeno-associated viruses (AAVs) are promising vectors in the field of gene therapy. Different AAV serotypes display distinct tissue tropism, believed to be related to the distribution of their receptors on target cells. Of the 11 well-characterized AAV serotypes, heparan sulfate proteoglycan and sialic acid have been suggested to be the attachment receptors for AAV type 2 and types 4 and 5, respectively. In this report, we identify the receptor for the two closely related serotypes, AAV1 and AAV6. First, we demonstrate using coinfection experiments and luciferase reporter analysis that AAV1 and AAV6 compete for similar receptors. Unlike heparin sulfate, enzymatic or genetic removal of sialic acid markedly reduced AAV1 and AAV6 binding and transduction. Further analysis using lectin staining and lectin competition assays identified that AAV1 and AAV6 use either alpha2,3-linked or alpha2,6-linked sialic acid when transducing numerous cell types (HepG2, Pro-5, and Cos-7). Treatment of cells with proteinase K but not glycolipid inhibitor reduced AAV1 and AAV6 infection, supporting the hypothesis that the sialic acid that facilitates infection is associated with glycoproteins rather than glycolipids. In addition, we determined by inhibitor (N-benzyl GalNAc)- and cell line-specific (Lec-1) studies that AAV1 and AAV6 require N-linked and not O-linked sialic acid. Furthermore, a resialylation experiment on a deficient Lec-2 cell line confirmed a 2,3 and 2,6 N-linked sialic acid requirement, while studies of mucin with O-linked sialic acid showed no inhibition effect for AAV1 and AAV6 transduction on Cos-7 cells. Finally, using a glycan array binding assay we determined that AAV1 efficiently binds to NeuAcalpha2-3GalNAcbeta1-4GlcNAc, as well as two glycoproteins with alpha2,3 and alpha2,6 N-linked sialic acids. Taken together, competition, genetic, inhibitor, enzymatic reconstitution, and glycan array experiments support alpha2,3 and alpha2,6 sialic acids that are present on N-linked glycoproteins as primary receptors for efficient AAV1 and AAV6 viral infection.  相似文献   

13.
A convenient chemoenzymatic strategy for synthesizing sialosides containing a C5-diversified sialic acid was developed. The α2,3- and α2,6-linked sialosides containing a 5-azido neuraminic acid synthesized by a highly efficient one-pot three-enzyme approach were converted to C5″-amino sialosides, which were used as common intermediates for chemical parallel synthesis to quickly generate a series of sialosides containing various sialic acid forms.  相似文献   

14.
Bronchial mucins from patients suffering from CF (cystic fibrosis) exhibit glycosylation alterations, especially increased amounts of the sialyl-Lewis(x) (NeuAcalpha2-3Galbeta1-4[Fucalpha1-3]GlcNAc-R) and 6-sulfo-sialyl-Lewis(x) (NeuAcalpha2-3Galbeta1-4[Fucalpha1-3][SO(3)H-6]GlcNAc-R) terminal structures. These epitopes are preferential receptors for Pseudomonas aeruginosa, the bacteria responsible for the chronicity of airway infection and involved in the morbidity and early death of CF patients. However, these glycosylation changes cannot be directly linked to defects in CFTR (CF transmembrane conductance regulator) gene expression since cells that secrete airway mucins express no or very low amounts of the protein. Several studies have shown that inflammation may affect glycosylation and sulfation of various glycoproteins, including mucins. In the present study, we show that incubation of macroscopically healthy fragments of human bronchial mucosa with IL-6 (interleukin-6) or IL-8 results in a significant increase in the expression of alpha1,3/4-fucosyltransferases [FUT11 (fucosyltransferase 11 gene) and FUT3], alpha2-6- and alpha2,3-sialyltransferases [ST3GAL6 (alpha2,3-sialyltransferase 6 gene) and ST6GAL2 (alpha2,6-sialyltransferase 2 gene)] and GlcNAc-6-O-sulfotransferases [CHST4 (carbohydrate sulfotransferase 4 gene) and CHST6] mRNA. In parallel, the amounts of sialyl-Lewis(x) and 6-sulfo-sialyl-Lewis(x) epitopes at the periphery of high-molecular-mass proteins, including MUC4, were also increased. In conclusion, our results indicate that IL-6 and -8 may contribute to the increased levels of sialyl-Lewis(x) and 6-sulfo-sialyl-Lewis(x) epitopes on human airway mucins from patients with CF.  相似文献   

15.
A Ca2+-independent sialic acid-specific lectin from two developmental stages of human placenta was similarly purified to apparent homogeneity by DEAE-cellulose chromatography, affinity chromatography on bovine submaxillary mucin, and gel filtration. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration disclosed a molecular mass of 53 kDa. The specificity of the lectin for O-acetylsialic acids was substantiated by the dependence of hemagglutination on the presence of acetylated sialic acids on the surface of mammalian erythrocytes of various sources, by hapten inhibition in hemagglutination assays with protease-treated rabbit erythrocytes and by hapten inhibition of binding of labeled N-acetylneuraminic acid-bovine serum albumin to the lectin in a solid-phase assay. Bovine and equine submaxillary mucins that contain 9(7,8)-O-acetyl and 4-O-acetylsialic acids were potent inhibitors in contrast to the non-acetylated sialic acids of ovine submaxillary mucin. Absence of inhibitory efficiency of other negatively charged substances like phosphorylated sugars, glucuronic acid, heparin, or oligodeoxynucleotides emphasized the importance of structural features instead of simple ionic interaction. In the presence of acetylation, the pattern of inhibition by gangliosides in the solid-phase assay indicated a preference to alpha-2,8- or alpha-2,6-linked sialic acids in comparison to alpha-2,3-linked moieties. Chemical modification of the lectin by group-specific reagents allowed to emphasize the role of primarily lysine residues, but also, although less pronounced, arginine, tryptophan, and carboxyl groups for ligand binding and/or maintenance of the active conformational state. Application of reagents, specific for histidine or tyrosine residues, failed to affect lectin activity.  相似文献   

16.
Group B streptococci (GBS) contain a capsular polysaccharide with side chains terminating in α2,3‐linked sialic acids. Because of this linkage type, the sialic acids of GBS are recognised by lectins of immune cells. This interaction results in a dampening of the host immune response and thus promotes immune evasion. As several influenza A viruses (IAV) use α2,3‐linked sialic acid as a receptor determinant for binding to host cells, we analysed whether GBS and influenza viruses can interact with each other and how this interaction affects viral replication and bacterial adherence to and invasion of host cells. A co‐sedimentation assay revealed that viruses with a preference for α2,3‐linked sialic acids bind to GBS in a sialic acid‐dependent manner. There is, however, a large variation in the efficiency of binding among avian influenza viruses of different subtypes as shown by a hemagglutination‐inhibition assay. A delay in the growth curve of IAV indicated that GBS has an inhibitory effect on virus replication. On the other hand, both the adherence and invasion efficiency of GBS were enhanced when the cells were pre‐infected by IAV with appropriate receptor specificity. Our results suggest that GBS infection may result in a more severe disease when patients are co‐infected by influenza viruses. This co‐infection mechanism may have relevance also to other human diseases, as there are more bacterial pathogens with α2,3‐linked sialic acids and human viruses binding to this linkage type.  相似文献   

17.
Avian influenza viruses of subtype H5N1 circulating in animals continue to pose threats to human health. The binding preference of the viral surface protein hemagglutinin (HA) to sialosaccharides of receptors is an important area for understanding mutations in the receptor binding site that could be the cause for avian-to-human transmission. In the present work, we studied the effect of two receptor binding site mutations, S221P singly and in combination with another mutation K216E in the HA protein of influenza A H5N1 viruses. Docking of sialic acid ligands corresponding to both avian and human receptors and molecular dynamics simulations of the complexes for wild and mutant strains of H5N1 viruses were carried out. The H5N1 strain possessing the S221P mutation indicated decreased binding to α2,3-linked sialic acids (avian receptor, SAα2,3Gal) when compared to the binding of the wild-type strain that did not possess the HA-221 mutation. The binding to α2,6-linked sialic acids (human receptor, SAα2,6Gal) was found to be comparable, indicating that the mutant strain shows limited dual receptor specificity. On the other hand, the S221P mutation in synergism with the K216E mutation in the binding site, resulted in increased binding affinity for SAα2,6Gal when compared to SAα2,3Gal, indicative of enhanced binding to human receptors. The in-depth study of the molecular interactions in the docked complexes could explain how co-occurring mutations in the HA viral protein can aid in providing fitness advantage to the virus, in the context of host receptor specificity in emerging variants of H5N1 influenza viruses.  相似文献   

18.
Expression of β1,6-branched N-linked oligosaccharides have a definite association with invasion and metastasis of cancer cells. However, the mechanism by which these oligosaccharides regulate these processes is not well understood. Invasive variants of B16 murine melanoma, B16F10 (parent) and B16BL6 (highly invasive variant) cell lines have been used for these studies. We demonstrate that substitution of α2,6-linked sialic acids on multiantennary structures formed as a result of β1,6-branching modulate cellular adhesion on both extracellular matrix (ECM) and basement membrane (BM) components. Removal of α2,6 sialic acids either by enzymatic desialylation or by stably down-regulating the ST6Gal-I (enzyme that catalyses the addition of α2,6-linked sialic acids on N-linked oligosaccharides) by lentiviral driven shRNA decreased the adhesion on both ECM and BM components and invasion through reconstituted BM matrigel.  相似文献   

19.
A series of dansyl-labeled glycosides with di-, tetra-, and hexasaccharides carrying the terminal N-acetyllactosamine (LacNAc) sequence were synthesized as acceptor substrates for α2,6- and α2,3-sialyltransferases. As an alternative design, dansyl-labeled LacNAc glycoside carrying a long-spacer linked glycan was engineered by replacement of the LacNAc or lactose units with an alkyl chain. In addition, we designed a dansyl-labeled bi-antennary LacNAc glycoside as an N-linked oligosaccharide mimetic, such as asialo-α(1)-acid glycoprotein. The kinetic parameters for the transfer reaction of synthesized dansyl-labeled glycosides by sialyltransferases were determined by the fluorescent HPLC method. The catalytic efficiencies (V(max)/K(m)) of acceptor substrates carrying the terminal LacNAc sequence with various length glycans in the array for α2,6- and α2,3-sialyltransferases decreased in a glycan length-dependent manner. Furthermore, of the acceptor substrates tested, dansyl-labeled bi-antennary LacNAc glycoside displayed the most favorable K(m) value for α2,6- and α2,3-sialyltransferases.  相似文献   

20.
The mouse-adapted A/PR/8/34 (PR8; H1N1) virus infects airway macrophages poorly and is virulent in mice. Herein, we have investigated factors contributing to the ability of PR8 to evade murine macrophages. We demonstrate that the hemagglutinin of PR8 binds preferentially to α(2,3)-linked sialic acid (SA) and that murine macrophages express α(2,6)-linked SA. Moreover, resialylation of macrophages to express α(2,3)-linked SA restored susceptibility to PR8. Thus, during adaptation of human influenza viruses to growth in mice, a switch in receptor specificity from α(2,6)-linked SA to α(2,3)-linked SA is likely to favour evasion of attachment, entry and destruction by airway macrophages.  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