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.
Serum alpha-fetoprotein (AFP) is a glycoprotein of which the sugar chain is considered to show structural changes with malignancies. Microheterogeneity of the serum AFP carbohydrate structure was studied in samples from 35 patients with benign and malignant diseases. Sera were digested directly, extensively, and sequentially with sialidase. beta-galactosidase and beta-N-acetylhexosaminidase. Before and after digestion, sera were examined by means of lectin affinity electrophoresis using eight lectins. Relationships between AFP carbohydrate structures and liver diseases were elucidated by the lectin-reactive profiles and the effect of glycosidase digestion. More than 94% of the AFP carbohydrate structures found in patients with benign and malignant liver diseases were biantennary complex-type oligosaccharides. Changes in the AFP carbohydrate structures at the early stage of hepatocellular carcinoma revealed the addition of alpha 1-->6 fucose to the reducing terminal N-acetylglucosamine and monosialylated AFPs. In both advanced hepatocellular carcinoma and AFP producing extrahepatic malignancies, AFP carbohydrate structures were characterized as the further addition of beta 1-->4 N-acetylglucosamine and heterogeneity in the galactose and N-acetylglucosamine residues. Sequential glycosidase digestion and lectin affinity electrophoresis is useful for analysing the carbohydrate structures of serum glycoprotein.  相似文献   

2.
Multimodal application of lectin affinity electrophoresis of alpha-fetoprotein (AFP) glycoforms is reviewed. Crossed affinity immunoelectrophoresis developed by B?g-Hansen and others was extended to tandem-lectin affinity electrophoresis by tandem lining of two different lectin gels and to mixed-lectin affinity electrophoresis. By introducing an antibody-affinity blotting technique for detection of separated glycoforms of AFP, several two-dimensional combinations of lectin affinity electrophoresis became possible: two different lectins for the first and second dimension electrophoresis, lectin-gradient affinity electrophoresis, and electrophoretic separation of lectin isoforms in the first-dimension electrophoresis, followed by affinity electrophoresis against the separated lectin isoforms. Usefulness of the different modalities of lectin affinity electrophoresis for several analytical purposes has been described.  相似文献   

3.
The lectins from the sponge Aaptos papillata were isolated by affinity chromatography using polyleucyl blood group A + H substances from hog stomach linings as an absorbent and eluting with 3 M MgCl2. Further separation on diethylaminoethylcellulose and preparative disc electrophoresis on polyacrylamide gave the three fractions, Aaptos lectins I, II, and III. They were essentially homogenous in polyacrylamide electrophoresis and sedimentation analysis: a small second component was seen in lectins I and II in immunoelectrophoresis at high concentration. The S20,W0 values for Aaptos lectins I, II, and III were 3.5, 6.0, and 5.5. By electrophoresis in sodium dodecyl sulfate with an without beta-mercaptoethanol Aaptos lectin I showed two bands corresponding to molecular weights of 12 000 and 21 000; Aaptos lectins II and III gave only one band of molecular weight of 16 000. In isoelectric focusing, Aaptos lectin I showed bands at pH 4.7 and 5.4 and in the range between 6.8 and 7.6, while Aaptos lectins II and III were almost identical with bands at pH 3.8, 4.7 to 4.9, and 5.3. Aaptos lectin I differed from II and III in amino acid composition but the latter two were very similar. They contained no significant carbohydrate. Aaptos lectin I reacted best with blood group substances with terminal nonreducing N-acetyl-D-glucosamine residues precipitating about two-thirds of the lectin N added while blood group substances with terminal nonreducing DGalNAc were almost inactive. However, Aaptos lectin II was completely precipitated by blood group substances and glycoproteins containing terminal DGalNAc, DGlcNAc, or sialic acid residues. Aaptos lectin III had a precipitation pattern similar to Aaptos lectin II. DGlcNAc but not DGalNAc inhibited precipitation of Aaptos lectin I by blood group substances and N, N', N', N'-tetraacetylchitotetraose was the best inhibitor and was 2000 times more active than DGlcNAc. Precipitin reactions with Aaptos lectin II were inhibited by equal amounts of DGlcNAc and by sialic acid which were four times more potent than DGalNAc. N,N',N'-triacetylchiotriose was the best inhibitor and was 13 times better than DGlcNAc. At 37 degrees C three to four times higher amounts of inhibitor were necessary to inhibit precipitation of Aaptos lectin II than were needed at 4 degrees C, indicating higher affinity of blood group substances for Aaptos lectin II with increasing temperature. Aaptos lectin I was precipitated by the monofunctional hapten p-nitrophenyl-alphaDGalNAc, while p-nitrophenyl-betaDGalNAc did not precipitate and was a good inhibitor. Both phenomena indicate involvement of hydrophobic bonds.  相似文献   

4.
The legume lectins are a large family of homologous carbohydrate binding proteins that are found mainly in the seeds of most legume plants. Despite their strong similarity on the level of their amino acid sequences and tertiary structures, their carbohydrate specificities and quaternary structures vary widely. In this review we will focus on the structural features of legume lectins and their complexes with carbohydrates. These will be discussed in the light of recent mutagenesis results when appropriate. Monosaccharide specificity seems to be achieved by the use of a conserved core of residues that hydrogen bond to the sugar, and a variable loop that determines the exact shape of the monosaccharide binding site. The higher affinity for particular oligosaccharides and monosaccharides containing a hydrophobic aglycon results mainly from a few distinct subsites next to the monosaccharide binding site. These subsites consist of a small number of variable residues and are found in both the mannose and galactose specificity groups. The quaternary structures of these proteins form the basis of a higher level of specificity, where the spacing between individual epitopes of multivalent carbohydrates becomes important. This results in homogeneous cross-linked lattices even in mixed precipitation systems, and is of relevance for their effects on the biological activities of cells such as mitogenic responses. Quaternary structure is also thought to play an important role in the high affinity interaction between some legume lectins and adenine and a series of adenine-derived plant hormones. The molecular basis of the variation in quaternary structure in this group of proteins is poorly understood.  相似文献   

5.
We studied the structure of N-linked carbohydrates bound to apolipoprotein H by a combination of two methods which make use of lectins. Digoxigenin-labelled lectins are used for the structural characterization of carbohydrate chains of glycoproteins. Concanavalin A lectin affinity chromatography was used to analyse apolipoprotein H according to the characteristics of its carbohydrate chain inner to sialic acid residues. Our results from digoxigenin-labelled lectins analysis showed that apolipoprotein H gave positive bands to SNA, DSA, GNA, PNA and AAA lectins. Apolipoprotein H gave a negative band when reacted with MAA lectin. When we applied apolipoprotein H onto the Concanavalin A lectin column no detectable amounts of protein were eluted with Concanavalin A buffer. After adding a buffer with low sugar concentration (10 mM glucoside) a large amount of apolipoprotein H was recovered. These molecules of apolipoprotein H weakly bound to the lectin. When a higher sugar concentration (500 mM mannoside) was added most of the sample applied was eluted. These molecules of apolipoprotein H firmly bound to the column having high affinity for the lectin. These results combined with those coming from the digoxigen-labeled lectins method enable us to understand the inner structure of carbohydrate chains with their outer branches. Molecules of apolipoprotein H which weakly bind to Concanavalin A could bear complex N-glycans organized in biantennary or truncated hybrid structures. Firmly bound apolipoprotein H referred to molecules rich in N-glycan hybrid structures. They have an outer branch belonging to the high mannose carbohydrate chains which explain the ability to bind to the column and an other main branch bearing the sequence galactose beta-(1-4)-N-acetylglucosamine beta-(1-2) mannose. Galactose could be the terminal sugar or, alternatively, be masked with sialic acid alpha-(2-6) terminally linked.  相似文献   

6.
The seed lectin from Dioclea grandiflora and jack bean lectin concanavalin A (ConA) are both members of the Diocleinae subtribe of Leguminosae lectins. Both lectins have recently been shown to possess enhanced affinities and extended binding sites for the trisaccharide, 3,6-di-O-(alpha-D-mannopyranosyl)-D-mannose, which is present in the core region of all asparagine-linked carbohydrates (Gupta, D., Oscarson, S., Raju, S., Stanley, P. Toone, E. J. and Brewer, C. F. (1996) Eur. J. Biochem. 242, 320-326). In the present study, the binding specificities of seven other lectins from the Diocleinae subtribe have been investigated by hemagglutination inhibition and isothermal titration microcalorimetry (ITC). The lectins are from Canavalia brasiliensis, Canavalia bonariensis, Cratylia floribunda, Dioclea rostrata, Dioclea virgata, Dioclea violacea, and Dioclea guianensis. Hemagglutination inhibition and ITC experiments show that all seven lectins are Man/Glc-specific and have high affinities for the core trimannoside, like ConA and D. grandiflora lectin. All seven lectins also exhibit the same pattern of binding to a series of monodeoxy analogs and a tetradeoxy analog of the trimannoside, similar to that of ConA and D. grandiflora lectin. However, C. bonariensis, C. floribunda, D. rostrata, and D. violacea, like D. grandiflora, show substantially reduced affinities for a biantennary complex carbohydrate with terminal GlcNAc residues, while C. brasiliensis, D. guianensis, and D. virgata, like ConA, exhibit affinities for the oligosaccharide comparable with that of the trimannoside. Thermodynamic data obtained by ITC indicate different energetic mechanisms of binding of the above two groups of lectins to the complex carbohydrate. The ability of the lectins to induce histamine release from rat peritoneal mast cells is shown to correlate with the relative affinities of the proteins for the biantennary carbohydrate.  相似文献   

7.
The carbohydrate residues of the surface coat of 20 axenic cultures of Blastocystis hominis were studied using FITC-labelled lectins (ConA, WGA, DBA, HPA, SBA, PNA, UEAI and LPA). The specific affinity of reactive lectins was determinated by competitive inhibition assay with specific carbohydrates or by enzymatic pre-treatment of cells. All stocks strongly bound ConA and HPA; WGA, UEAI and LPA were partially reactive, and the remaining lectins were nonreactive. Inhibition assays showed abolition (WGA, LPA, UEAI and HPA) or partial reduction (ConA) of lectin affinity, which demonstrated the specificity of binding assay. These results indicate that B. hominis has surface components containing alpha-D-mannose, alpha-D-glucose, N-acetyl-alpha-D-glucosamine, alpha-L-fucose, chitin and sialic acid.  相似文献   

8.
Both the Entamoeba histolytica lectin, a virulence factor for the causative agent of amebiasis, and the mammalian hepatic lectin bind to N-acetylgalactosamine (GalNAc) and galactose (Gal) nonreducing termini on oligosaccharides, with preference for GalNAc. Polyvalent GalNAc-derivatized neoglycoproteins have >1000-fold enhanced binding affinity for both lectins (Adler,P., Wood,S.J., Lee,Y.C., Lee,R.T., Petri,W.A.,Jr. and Schnaar,R.L.,1995, J. Biol. Chem ., 270, 5164-5171). Substructural specificity studies revealed that the 3-OH and 4-OH groups of GalNAc were required for binding to both lectins, whereas only the E.histolytica lectin required the 6-OH group. Whereas GalNAc binds with 4-fold lower affinity to the E.histolytica lectin than to the mammalian hepatic lectin, galactosamine and N-benzoyl galactosamine bind with higher affinity to the E. histolytica lectin. Therefore, a synthetic scheme for converting polyamine carriers to poly-N-acyl galactosamine derivatives (linked through the galactosamine primary amino group) was developed to test whether such ligands would bind the E.histolytica lectin with high specificity and high affinity. Contrary to expectations, polyvalent derivatives including GalN6lys5, GalN4desmosine, GalN4StarburstTMdendrimer, and GalN8StarburstTMdendrimer demonstrated highly enhanced binding to the mammalian hepatic lectin but little or no enhancement of binding to the E.histolytica lectin. We propose that the mammalian hepatic lectin binds with greatest affinity to GalNAc "miniclusters," which mimic branched termini of N-linked oligosaccharides, whereas the E.histolytica lectin binds most effectively to "maxiclusters," which may mimic more widely spaced GalNAc residues on intestinal mucins.  相似文献   

9.
Calnexin and calreticulin are homologous molecular chaperones of the endoplasmic reticulum. Their binding to newly synthesized glycoproteins is mediated, at least in part, by a lectin site that recognizes the early N-linked oligosaccharide processing intermediate, Glc1Man9GlcNAc2. We compared the oligosaccharide binding specificities of calnexin and calreticulin in an effort to determine the basis for reported differences in their association with various glycoproteins. Using mono-, di-, and oligosaccharides to inhibit the binding of Glc1Man9GlcNAc2 to calreticulin and to a truncated, soluble form of calnexin, we show that the entire Glc alpha 1-3Man alpha 1-2Man alpha 1-2Man structure, extending from the alpha 1-3 branch point of the oligosaccharide core, is recognized by both proteins. Furthermore, analysis of the binding of monoglucosylated oligosaccharides containing progressively fewer mannose residues suggests that for both proteins the alpha 1-6 mannose branch point of the oligosaccharide core is also essential for recognition. Consistent with their essentially identical substrate specificities, calnexin and calreticulin exhibited the same relative affinities when competing for binding to the Glc1Man9GlcNAc2 oligosaccharide. Thus, differential glycoprotein binding cannot be attributed to differences in the lectin specificities or binding affinities of calnexin and calreticulin. We also examined the effects of ATP, calcium, and disulfide reduction on the lectin properties of calnexin and calreticulin. Whereas oligosaccharide binding was only slightly enhanced for both proteins in the presence of high concentrations of a number of adenosine nucleotides, removal of bound calcium abrogated oligosaccharide binding, an effect that was largely reversible upon readdition of calcium. Disulfide reduction had no effect on oligosaccharide binding by calnexin, but binding by calreticulin was inhibited by 70%. Finally, deletion mutagenesis of calnexin and calreticulin identified a central proline-rich region characterized by two tandem repeat motifs as a segment capable of binding oligosaccharide. This segment bears no sequence homology to the carbohydrate recognition domains of other lectins.  相似文献   

10.
By performing electrophoresis perpendicular to a stationary pH gradient (pH-mobility curves) in polyacrylamide gels containing a specific ligand either covalently fixed or entrapped in the gel matrix, it is possible to measure dissociation constants (Kd) and their pH-dependence in the pH range 3.5-10. The present technique, called 'affinity titration curves', is an extension of 'affinity electrophoresis'. This system has been applied to the study of the interaction between lectins and sugars: lectin from Ricinus communis seeds and alpha-D-galactose, and lectin from Lens culinaris seeds and alpha-D-mannose. The pH-dependence of Kd values indicated a more rapid decrement of affinity of both lectins for their ligands at acidic pH as compared to alkaline pH. For both lectins, maximum affinity was found in the pH range 7-8. Since the ionic strength of focused carrier ampholytes is 100-200-times lower than in conventional electrophoresis, the Kd values found by the present method are generally lower than the same values obtained by affinity electrophoresis.  相似文献   

11.
Purified porcine thyroglobulin (Tg) was fractionated on a concanavalin A-Sepharose 4B column by a step-wise elution with increasing concentrations of methyl alpha-mannoside (fraction A, 50 mM; B, 100 mM; C, 200 mM; D, 500 mM, and E, 1 M), and its fractional ratio was 12.8:28.6:26.4:19.7:12.4. These five fractions showed the same profile in polyacrylamide gel electrophoresis. The subfractions were analyzed for their relative contents in oligosaccharides of each structure type and for their monosaccharide contents. In fractions B, C, D, and E the former varied between 15-22% for triantennary complex-type, 47-60% for biantennary complex-type, and 22-30% for high mannose-type oligosaccharide. Fraction A showed a higher percentage of triantennary complex-type structures (36%) and a lower percentage of biantennary complex-type structures (17%). The monosaccharide numbers increased from fraction A to E: 85 to 135 mannose residues, 60 to 82 galactose residues, 84 to 115 N-acetyl glucosamine residues, and 22 to 28 sialic acid residues. After analysis of the number of mannose residues contained in the high mannose-type structures, it was possible to calculate the number of oligosaccharides borne by each Tg subfraction. This number was approximately the same for fractions A and B (22.4 and 21.7), then it increased from B to E (21.8 to 32.9). These results account for the separation obtained on the concanavalin A-Sepharose 4B column. Separation of the two first subfractions bearing the same number of oligosaccharides is certainly due to the higher number of high mannose-type structures in B. In conclusion, the studies reported here show that porcine Tg is heterogeneous, and mainly so in terms of total number of N-glycan structures.  相似文献   

12.
The synthesis and binding properties of a new family of high affinity alpha-D-mannopyranoside ligands are described. The synthesis of the new multivalent ligands is based on the scaffolding of multiantennary branches of L-lysine residues having electrophilic N-chloroacetylated end groups as core structures. An alpha-D-mannopyranoside with p-substituted aryl aglycon ending with a thiol group was prepared and covalently attached to each of the branches of the dendritic structures. The resulting glycodendrimers with 2 (12), 4 (14), 8 (16), and 16 (18) mannoside residues were tested for their relative inhibitory potency by solid-phase enzyme-linked lectin assays (ELLA) using methyl and p-nitrophenyl alpha-D-mannopyranosides as standards. Concentrations necessary for 50% inhibition (IC50s) of binding of yeast mannan to Jack bean phytohemagglutinin (Canavalia ensiformis, concanavalin A) and to pea lectin (Pisum sativum) were determined. Analogous mannosylated copolyacrylamides were also prepared for comparison. The IC50 values were also plotted as a function of dendrimer valencies. The inhibitions showed 16-mer 18 to be approximately 600- and 2000-fold more potent than methyl alpha-D-mannopyranoside, and 66- and 1383-fold more potent than p-nitrophenyl alpha-D-mannopyranosides with Con A and pea lectins, respectively. Even when these numbers are expressed relative to single mannopyranoside residues per dendrimers, the relative potencies against the aromatic mannoside are still 4- and 86-fold better against Con A and pea lectins. These results unequivocally indicate that the optimum inhibitory binding properties of the new mannosylated dendrimers vary with both dendrimers and lectin valencies.  相似文献   

13.
Glycoforms of recombinant human follicle stimulating hormone (rhFSH) (Org 32489, Puregon) were characterized using concanavalin A lectin affinity chromatography to reveal information about the internal carbohydrate complexity (extent of carbohydrate side-chain branching) of the preparations. The rhFSH glycoforms were measured by radioimmunoassay and a two-site immunoradiometric assay and compared with those in two urinary preparations (Metrodin and Metrodin-HP) used in assisted reproduction programmes and a urinary FSH international standard 70/45 (uFSH IS 70/45). Similar data were obtained with both assays; rhFSH had 6% complex internal carbohydrate structures compared with 22-27% for Metrodin, Metrodin-HP and uFSH. The proportion of simple carbohydrate structures was also different, with rhFSH having 18.5 compared with 4.5-9.3% for Metrodin, Metrodin-HP and uFSH. A linear relationship was observed between the percentage glycoforms with an isoelectric point (pl) < 4 and the log percentage simple forms (logarithmic regression; r = 0.93) indicating a direct relationship between carbohydrate complexity and charge heterogeneity. In summary, rhFSH contains fewer complex forms and an increased proportion of simple carbohydrate structures in comparison with Metrodin, Metrodin-HP and IS 70/45.  相似文献   

14.
The hemagglutinins from the sponge Axinella polypoides were isolated by affinity chromatography using Sepharose 4B as an absorbent and eluting with DGal. Further separation on DEAE-cellulose and preparative disc electrophoresis on polyacrylamide and agarose gave three fractions. The physicochemical properties and binding specificities of the two main agglutinins were studied. Homogeneity was tested by polyacrylamide electrophoresis and immunoelectrophoresis and by sedimentation analysis. In isoelectric focusing, agglutinin I (mol wt 21 000) showed two bands at pH 3.8 and 3.9. Agglutinin II (mol wt 15 000) showed one band at pH 3.9. Both agglutinins have a carbohydrate content of about 0.5%, are immunochemically unrelated, and differ in amino acid composition. Both precipitate A1, A2, B, Lea, and precursor I blood group substances but to different extents. Inhibition experiments revealed that both agglutinins are inhibited best by terminal nonreducing DGal glycosidically linked beta 1 leads to 6 or by p-nitrophenyl-betaDGal. DGal and DFuc are equally active but about 20 and 12 times less active with agglutinin I and agglutinin II, respectively. DGalNAc and LFuc were inactive even at much higher concentrations. Both agglutinins have similar specificities and react with the immunodominant determinants of blood group B and Lea but not with A and H substances; in A and H substances, reactivity is with side chains in which beta-linked DGal is unsubstituted at the nonreducing terminus. The Axinella polypoides lectins are compared with galactose-specific lectins of different origin and with the aggregation factor system is sponges.  相似文献   

15.
The seed lectin from Dioclea grandiflora (DGL) has recently been shown to possess high affinity for 3, 6-di-O-(alpha-D-mannopyranosyl)-alpha-D-mannopyranose, the core trimannoside of asparagine-linked carbohydrates, but lower affinity for biantennary complex carbohydrates. In the previous paper, the thermodynamics of DGL binding to deoxy analogs of the core trimannoside and to a biantennary complex carbohydrate were determined by isothermal titration microcalorimetry. The data suggest that DGL recognizes specific hydroxyl groups of the trimannoside similar to that of the jack bean lectin concanavalin A (ConA) (Gupta, D. Dam, T. K., Oscarson, S., and Brewer, C. F. (1997) J. Biol. Chem. 272, 6388-6392). However, the thermodynamics of DGL binding to certain deoxy analogs and to the complex carbohydrate are different from that of ConA. In the present paper, the x-ray crystal structure of DGL complexed to the core trimannoside was determined to a resolution of 2.6 A. The overall structure of the DGL complex is similar to the structure of the ConA-trimannoside complex (Naismith, J. H., and Field, R. A. (1996) J. Biol. Chem. 271, 972-976). The location and conformation of the bound trimannoside as well as its hydrogen-bonding interactions in both complexes are nearly identical. However, differences exist in the location of two loops outside of the respective binding sites containing residues 114-125 and 222-227. The latter residues affect the location of a network of hydrogen-bonded water molecules that interact with the trisaccharide. Differences in the arrangement of ordered water molecules in the binding site and/or protein conformational differences outside of the binding site may account for the differences in the thermodynamics of binding of the two lectins to deoxy analogs of the trimannoside. Molecular modeling studies suggest how DGL discriminates against binding the biantennary complex carbohydrate relative to ConA.  相似文献   

16.
Two mannose-binding lectins, Allium sativum agglutinin (ASA) I (25 kDa) and ASAIII (48 kDa), from garlic bulbs have been purified by affinity chromatography followed by gel filtration. The subunit structures of these lectins are different, but they display similar sugar specificities. Both ASAI and ASAIII are made up of 12.5- and 11.5-kDa subunits. In addition, a complex (136 kDa) comprising a polypeptide chain of 54 +/- 4 kDa and the subunits of ASAI and ASAIII elutes earlier than these lectins on gel filtration. The 54-kDa subunit is proven to be alliinase, which is known to form a complex with garlic lectins. Constituent subunits of ASAI and ASAIII exhibit the same sequence at their amino termini. ASAI and ASAIII recognize monosaccharides in mannosyl configuration. The potencies of the ligands for ASAs increase in the following order: mannobiose (Manalpha1-3Man) < mannotriose (Manalpha1-6Manalpha1-3Man) approximately mannopentaose < Man9-oligosaccharide. The addition of two GlcNAc residues at the reducing end of mannotriose or mannopentaose enhances their potencies significantly, whereas substitution of both alpha1-3- and alpha1-6-mannosyl residues of mannotriose with GlcNAc at the nonreducing end increases their activity only marginally. The best manno-oligosaccharide ligand is Man9GlcNAc2Asn, which bears several alpha1-2-linked mannose residues. Interaction with glycoproteins suggests that these lectins recognize internal mannose as well as bind to the core pentasaccharide of N-linked glycans even when it is sialylated. The strongest inhibitors are the high mannose-containing glycoproteins, which carry larger glycan chains. Indeed, invertase, which contains 85% of its mannose residues in species larger than Man20GlcNAc, exhibited the highest binding affinity. No other mannose- or mannose/glucose-binding lectin has been shown to display such a specificity.  相似文献   

17.
Artocarpin, a mannose-specific lectin, is a homotetrameric protein (M(r) 65,000) devoid of covalently attached carbohydrates and consists of four isolectins with pI in the range 5-6.5. Investigations of its carbohydrate binding specificity reveal that among monosaccharides, mannose is preferred over glucose. Among mannooligosaccharides, mannotriose (Man alpha 1-3[Man alpha 1-6]Man) and mannopentaose are the strongest ligands followed by Man alpha 1-3Man. Extension of these ligands by GlcNAc at the reducing ends of mannooligosaccharides tested remarkably improves their inhibitory potencies, while substitution of both the alpha 1-3 and alpha 1-6 mannosyl residues of mannotriose and the core pentasaccharide of N-linked glycans (Man alpha 1-3[Man alpha 1-6]Man beta 1-4GlcNAc beta 1-4GlcNAc) by GlcNAc or N-acetyllactosamine in beta 1-2 linkage diminishes their inhibitory potencies. Sialylated oligosaccharides are non-inhibitory. Moreover, the substitution of either alpha 1-3 or alpha 1-6 linked mannosyl residues of M5Gn or both by mannose in alpha 1-2 linkage leads to a considerable reduction of their inhibitory power. Addition of a xylose residue in beta 1-2 linkage to the core pentasaccharide improves the inhibitory activity. Considering the fact that artocarpin has the strongest affinity for the xylose containing hepasaccharide from horseradish peroxidase, which differs significantly from all the mannose/glucose-specific lectins, it should prove a useful tool for the isolation and characterization of glycoproteins displaying such structure.  相似文献   

18.
The sugar residues in glycoconjugates present in the parotid and mandibular glands of the adult fallow-deer were detected and characterized by using a battery of eight different lectin-horseradish peroxidase conjugates. In some cases a treatment with sialidase preceded the lectin staining. Parotid secretory cells produced glycoconjugates with N-acetylgalactosamine, N-acetylglucosamine and mannose residues. Mucous acinar cells were the most reactive sites of the mandibular gland and contained conspicuous quantities of oligosaccharides with terminal sialic acid radicals. Galactosil-(beta 1-->3)N-acetylgalactosamine was the most abundant penultimate sugar linked to N-acetylneuraminic acid. Mandibular mucous cells also presented N-acetylglucosamine and sialylated components with the terminal dimer sialic acid-N-acetylgalactosamine. Demilunar cells contained glycoconjugates with fucose and mannose residues. The apical surface of duct cells was stained by all the lectins.  相似文献   

19.
Clusters of di-, tri-, and tetra-antennary alpha-D-mannopyranosides were synthesized in good yields based on the coupling of amine-bearing mono- or trisaccharide [Man alpha(1 --> 6)[Man alpha(1 --> 3)]Man] haptens to poly-isocyanate or -isothiocyanate tethering cores. The relative binding properties of the resulting multivalent ligands were determined by turbidimetric and solid phase enzyme-linked lectin assays (ELLA) using plant lectins (phytohemagglutinins) Concanavalin A (Con A) and Pisum sativum (pea lectin) having four and two carbohydrate binding sites, respectively. Rapid and efficient cross-linking between tetravalent Con A and mannopyranosylated clusters were measured by a microtiter plate version of turbidimetric analyses. In inhibition of binding of the lectins to yeast mannan, the best tetravalent monosaccharide (30) and trisaccharide (31) inhibitors were shown to be 140 and 1155 times more potent inhibitors than monomeric methyl alpha-D-mannopyranoside against pea lectin and Con A, respectively. Compounds 30 and 31 were thus 35- and 289-fold more potent than the reference monosaccharide based on their hapten contents. As a general observation, the ligands bearing the Man alpha(1 --> 6)[Man alpha(1 --> 3)]Man trimannoside structures were found to be more potent inhibitors for Con A than the ligands having single mannoside residues, whereas pea lectin could not discriminate between the two types of ligands.  相似文献   

20.
The type II antifreeze proteins (AFPs) of smelt and Atlantic herring are homologous to the carbohydrate-recognition domains (CRDs) of Ca2+-dependent (C-type) animal lectins and, like these lectins, acquire a stable and active structure upon binding Ca2+ ions. In the C-type lectin CRD, the carbohydrate-binding site is located at a Ca2+-binding site. Site-directed mutagenesis was used to test the hypothesis that the ice-binding site of the type II AFP corresponds to the carbohydrate-binding site of the lectins. To disrupt this site in the herring AFP without perturbing the Ca2+-dependent protein fold, a double mutant was constructed that changed the Ca2+- and carbohydrate-binding motif from the galactose-type of wild-type AFP containing the sequence Gln-Pro-Asp to a mannose-type that has the sequence Glu-Pro-Asn and is also known to bind Ca2+. The mutant AFP exhibited proper Ca2+ binding, folding, and stability as demonstrated by ruthenium red staining, proteolysis protection assays, and CD spectroscopy. However, it showed no antifreeze activity (thermal hysteresis) and did not alter ice crystal morphology to form bipyramidal crystals as does the active wild-type AFP. These results demonstrate that the ice-binding site of the herring type II AFP corresponds to the carbohydrate-binding site of the C-type lectin CRDs and further suggest that this ice-binding function evolved from the carbohydrate-binding site of a preexisting C-type lectin.  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