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.
A hallmark of oligosaccharides is their often limited spatial flexibility, allowing them to access a distinct set of conformers in solution. Viewing each individual or even the complete ensemble of conformations as potential binding partner(s) for lectins in protein–carbohydrate interactions, it is pertinent to address the question on the characteristics of bound state conformation(s) in solution. Also, it is possible that entering the lectin’s binding site distorts the low-energy topology of a glycosidic linkage. As a step to delineate the strategy of ligand selection for galactosides, a common physiological docking point, we have performed a NMR study on two non-homologous lectins showing identical monosaccharide specificity. Thus, the conformation of lactose analogues bound to bovine heart galectin-1 and to mistletoe lectin in solution has been determined by transferred nuclear Overhauser effect measurements. It is demonstrated that the lectins select the syn conformation of lactose and various structural analogues (Galβ(1→4)Xyl, Galβ(1→3)Xyl, Galβ(1→2)Xyl, and Galβ(1→3)Glc) from the ensemble of presented conformations. No evidence for conformational distortion was obtained. Docking of the analogues to the modeled binding sites furnishes explanations, in structural terms, for exclusive recognition of the syn conformer despite the non-homologous design of the binding sites.  相似文献   

2.
Glycosphingolipid recognition by two isolectins from Solanum tuberosum was compared by the chromatogram binding assay. One lectin (PL-I) was isolated from potato tubers by affinity chromatography, and identified by MALDI-TOF mass spectrometry as a homodimer with a subunit molecular mass of 63,000. The other (PL-II) was a commercial lectin, characterized as two homodimeric isolectins with subunit molecular masses of 52,000 and 55,000, respectively. Both lectins recognized N-acetyllactosamine-containing glycosphingolipids, but the fine details of their carbohydrate binding specificities differed. PL-II preferentially bound to glycosphingolipids with N-acetyllactosamine branches, as Galbeta4GlcNAcbeta6(Galbeta4GlcNAcbeta3)Galbeta4Glcbeta1C er. PL-I also recognized this glycosphingolipid, but bound equally well to the linear glycosphingolipid Galbeta4GlcNAcbeta3Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer. Neolactotetraosylceramide and the B5 pentaglycosylceramide were also bound by PL-I, while other glycosphingolipids with only one N-acetyllactosamine unit were non-binding. Surprisingly, both lectins also bound to lactosylceramide, with an absolute requirement for sphingosine and non-hydroxy fatty acids. The inhibition of binding to both lactosylceramide and N-acetyllactosamine-containing glycosphingolipids by N-acetylchitotetraose suggests that lactosylceramide is also accomodated within the N-acetylchitotetraose/N-acetyllactosamine-binding sites of the lectins. Through docking of glycosphingolipids onto a three-dimensional model of the PL-I hevein binding domain, a Galbeta4GlcNAcbeta3Galbeta4 binding epitope was defined. Furthermore, direct involvement of the ceramide in the binding of lactosylceramide was suggested.  相似文献   

3.
Galbeta1-3GalNAc (T-disaccharide) and related molecules were assayed to describe the structural requirements of carbohydrates to bind Agaricus bisporus lectin (ABL). Results provide insight into the most relevant regions of T-disaccharide involved in the binding of ABL. It was found that monosaccharides bind ABL weakly indicating a more extended carbohydrate-binding site as compared to those involvedin the T- disaccharide specific lectins such as jacalin and peanut agglutinin. Lacto-N-biose (Galbeta1-3GlcNAc) unlike T-disaccharide, is unable to inhibit the ABL interaction, thus showing the great importance of the position of the axial C-4 hydroxyl group of GalNAc in T-disaccharide. This finding could explain the inhibitory ability of Galbeta1-6GlcNAc and lactose because C-4 and C-3 hydroxyl groups of reducing Glc, respectively, occupy a similar position as reported by conformational analysis. From the comparison of different glycolipids bearing terminal T-disaccharide bound to different linkages, it can be seen than ABL binding is even more impaired by an adjacent C-6 residual position than by the anomeric influence of T-disaccharide. Furthermore, the addition of beta-GlcNAc to the terminal T-disaccharide in C-3 position of Gal does not affect the ABL binding whereas if an anionic group such as glucuronic acid is added to C-3, the binding is partially affected. These findings demonstrate that ABL holds a particular binding nature different from that of other T-disaccharide specific lectins.   相似文献   

4.
Lycknert K  Edblad M  Imberty A  Widmalm G 《Biochemistry》2004,43(30):9647-9654
The beta-D-GlcpNAc-(1-->6)-alpha-D-Manp disaccharide is a constituent of highly branched cell-surface glycoconjugates that are malignancy markers. The conformational preference of the disaccharide beta-D-GlcpNAc-(1-->6)-alpha-D-Manp-OMe in solution has been studied by molecular modeling and NMR spectroscopy including 1D (1)H,(1)H T-ROESY experiments and analysis of (3)J(H,H) of the hydroxymethyl group being part of the glycosidic linkage of the disaccharide, which revealed the relative populations of the omega torsion angle as gt = 0.60, gg = 0.35, and tg = 0.05. Good agreement was obtained between the effective proton-proton distances from the experiment and those obtained by molecular modeling when the flexibility at the omega torsion angle was taken into account. Molecular modeling of the disaccharide in the binding sites of the lectin wheat germ agglutinin indicates that several conformations could be adopted in the bound state. (1)H NMR and transfer NOESY experiments confirmed that binding took place, and trans-glycosidic proton-proton interactions indicated that a conformational preference was present in the bound state, as observed by the relative change of the NOEs from H1' to H6(pro-R) and H6(pro-S). STD NMR experiments showed that binding occurred in the region of the N-acetyl group of the terminal sugar residue. In addition, the O-methyl group received saturation transfer because of the proximity to the protein. (1)H,(1)H NOEs indicated that the two methyl groups were close in space, as observed in only one of the predicted bound conformations. Experimental and theoretical data therefore agree that one conformation with a gt conformation of the hydroxymethyl group and a negative sign for the psi torsion angle is indeed selected by the lectin upon binding.  相似文献   

5.
G F Leanz  G G Hammes 《Biochemistry》1986,25(19):5617-5624
The ionic strength dependence of the second-order rate constant for the association of reduced nicotinamide adenine dinucleotide phosphate (NADPH) and chicken liver fatty acid synthase was determined. This rate constant is 7.2 X 10(7) M-1 s-1 at zero ionic strength and 25 degrees C; the effective charge at the cofactor binding sites is +0.8. The conformations of nicotinamide adenine dinucleotide phosphate (NADP+) and NADPH bound to the beta-ketoacyl and enoyl reductase sites were determined from transferred nuclear Overhauser effect measurements. Covalent modification of the enzyme with pyridoxal 5'-phosphate abolished cofactor binding at the enoyl reductase site; this permitted the cofactor conformations at the beta-ketoacyl and enoyl reductase sites to be distinguished. For NADP+ bound to the enzyme, the conformation of the nicotinamide-ribose bond is anti at the enoyl reductase site and syn at the beta-ketoacyl reductase site; the adenine-ribose bond is anti, and the sugar puckers are C3'-endo. Nicotinamide-adenine base stacking was not detected. Structural models of NADP+ at the beta-ketoacyl and enoyl reductase sites were constructed by using the distances calculated from the observed nuclear Overhauser effects. Because of the overlap of the resonances of several nonaromatic NADPH protons with the resonances of HDO and ribose protons, less extensive structural information was obtained for NADPH bound to the enzyme. However, the conformations of NADPH bound to the two reductases are qualitatively the same as those of NADP+, except that the nicotinamide moiety of NADPH is closer to being fully anti at the enoyl reductase site.  相似文献   

6.
A model of the carbohydrate recognition domain of the serum form of mannose-binding protein (MBP) from rat complexed with methyl 3,6-di-O-(alpha-D-mannopyranosyl)-alpha-D-mannopyranoside is presented. Allowed conformations for the bound sugar were derived from simulated annealing protocols incorporating distance restraints computed from transferred NOESY spectra. The resulting sugar conformations were then modeled into the MBP binding site, and these models of the complex were refined using molecular dynamics (MD) simulations in the presence of solvent water. These studies indicate that only one of the two major conformations of the alpha(1-->6) linkage found in solution is significantly populated in the bound state (omega = 60 degrees ), whereas the alpha(1-->3) linkage samples at least two states, similar to its behavior in free solution. The bound conformation allows direct hydrogen bonds to form between the sugar and K182 of MBP, in addition to other water-mediated hydrogen bonds. Estimates of binding constants of candidate complexes based on changes in solvent-accessible surface areas upon binding support the NMR and MD results. These estimates further suggest that the enthalpic gains of the additional sugar-MBP interactions in a trisaccharide as opposed to a monosaccharide are offset by entropic penalties, offering an explanation for previous binding data.  相似文献   

7.
P R Rosevear  T L Fox  A S Mildvan 《Biochemistry》1987,26(12):3487-3493
MgATP binds both at the active site (site 1) and at a secondary site (site 2) on each monomer of muscle pyruvate kinase as previously found by binding studies and by X-ray analysis. Interproton distances on MgATP bound at each site have been measured by the time-dependent nuclear Overhauser effect in the absence and presence of phosphoenolpyruvate (P-enolpyruvate), which blocks ATP binding at site 1. Interproton distances at site 2 are consistent with a single conformation of bound ATP with a high antiglycosidic torsional angle (chi = 68 +/- 10 degrees) and a C3'-endo ribose pucker (delta = 90 +/- 10 degrees). Interproton distances at site 1, determined in the absence of P-enolpyruvate by assuming the averaging of distances at both sites, cannot be fit by a single adenine-ribose conformation but require the contribution of at least three low-energy structures: 62 +/- 10% low anti (chi = 30 degrees), C3'-endo; 20 +/- 8% high anti (chi = 55 degrees), O1'-endo; and 18 +/- 8% syn (chi = 217 degrees), C2'-endo. Although a different set of ATP conformations might also have fit the interproton distances, the mixture of conformations used also fits previously determined distances from Mn2+ to the protons of ATP bound at site 1 [Sloan, D. L., & Mildvan, A. S. (1976) J. Biol. Chem. 251, 2412] and is similar to the adenine-ribose portion of free Co(NH3)4ATP, which consists of 35% low anti, 51% high anti, and 14% syn [Rosevear, P. R., Bramson, H. N., O'Brian, C., Kaiser, E. T., & Mildvan, A. S. (1983) Biochemistry 22, 3439].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The long-held views on lock-and-key versus induced fit in binding arose from the notion that a protein exists in a single, most stable conformation, dictated by its sequence. However, in solution proteins exist in a range of conformations, which may be described by statistical mechanical laws and their populations follow statistical distributions. Upon binding, the equilibrium will shift in favor of the bound conformation from the ensemble of conformations around the bottom of the folding funnel. Hence here we extend the implications and the usefulness of the folding funnel concept to explain fundamental binding mechanisms.  相似文献   

9.
The carbohydrate binding preferences of the Galalpha3Galbeta4 GlcNAc-binding lectins from Marasmius oreades and Euonymus europaeus were examined by binding to glycosphingolipids on thin-layer chromatograms and in microtiter wells. The M. oreades lectin bound to Galalpha3-terminated glycosphingolipids with a preference for type 2 chains. The B6 type 2 glycosphingolipid (Galalpha3[Fucalpha2]Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer) was preferred over the B5 glycosphingolipid (Galalpha3Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer), suggesting that the alpha2-linked Fuc is accommodated in the carbohydrate binding site, providing additional interactions. The lectin from E. europaeus had broader binding specificity. The B6 type 2 glycosphingolipid was the best ligand also for this lectin, but binding to the B6 type 1 glycosphingolipid (Galalpha3[Fucalpha2]Galbeta3GlcNAcbeta3Galbeta4Glcbeta1Cer) was also obtained. Furthermore, the H5 type 2 glycosphingolipid (Fucalpha2Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer), devoid of a terminal alpha3-linked Gal, was preferred over the the B5 glycosphingolipid, demonstrating a significant contribution to the binding affinity by the alpha2-linked Fuc. The more tolerant nature of the lectin from E. europaeus was also demonstrated by the binding of this lectin, but not the M. oreades lectin, to the x2 glycosphingolipid (GalNAcbeta3Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer) and GlcNAcbeta3Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer. The A6 type 2 glycosphingolipid (GalNAcalpha3[Fucalpha2]Galbeta4GlcNAcbeta3Galbeta4Glcbeta1Cer) and GalNAcalpha3Galbeta4GlcNAcbeta3Galbeta4Glcbeta1-Cer were not recognized by the lectins despite the interaction with B6 type 2 glycosphingolipid and the B5 glycosphingolipid. These observations are explained by the absolute requirement of a free hydroxyl in the 2-position of Galalpha3 and that the E. europaea lectin can accommodate a GlcNAc acetamido moiety close to this position by reorienting the terminal sugar, whereas the M. oreades lectin cannot.  相似文献   

10.
Kinetic parameters of the interaction of the toxic lectins abrin and ricin with human erythrocytes and HeLa cells have been measured. The binding of 125I-labeled abrin and ricin to human erythrocytes and to HeLa cells at 37 degrees was maximal around pH 7, whereas at 0 degrees the binding was similar over a broad pH range. The binding occurred at similar rates at 0 degrees and 37 degrees with rate constants in the range 0.9 to 3.0 X 10(5) M-1 s-1. The dissociation was strongly temperature-dependent with rate constants in the range 3.4 to 45 X 10(-4) s-1 at 0 degrees and 3.9 to 18 X 10(-3) s-1 at 37 degrees. The presence of unlabeled lectins as well as lactose increased the rate of dissociation. The association constants measured at equilibrium or calculated from the rate constants were between 0.64 X 10(8) M-1 and 8.2 X 10(8) M-1 for abrus lectins, and between 8.0 X 10(6) M-1 and 4.2 X 10(8) M-1 for ricinus lectins. The association constants for the toxins were lower at 37 degrees than at 0 degrees. Isolated ricin B chain appeared to bind with similar affinity as intact ricin. The number of binding sites was estimated to be 2 to 3 X 10(6) per erythrocyte and 1 to 3 X 10(7) per HeLa cell. The binding sites of HeLa cells all displayed a uniform affinity towards abrin and ricin, both at 0 degrees and at 37 degrees. The same was the case with the binding sites of erythrocytes at 0 degrees. However, the data indicated that at 20 degrees erythrocytes possessed binding sites with two different affinities. Only a fraction of the cell-bound toxin appeared to be irreversibly bound and could not be removed by washing with 0.1 M lactose. The fraction of the total amount of bound toxin which became irreversibly bound to HeLa cells was for both toxins about 2 X 10(-3)/min at 37 degrees, whereas no toxin was irreversibly bound at 0 degrees. In the case of erythrocytes no toxin became irreversibly bound, either at 0 degrees or 37 degrees, indicating that the toxins are unable to penetrate into these cells.  相似文献   

11.
We employ ensemble docking simulations to characterize the interactions of two enantiomeric forms of a Ru-complex compound (1-R and 1-S) with three protein kinases, namely PIM1, GSK-3β, and CDK2/cyclin A. We show that our ensemble docking computational protocol adequately models the structural features of these interactions and discriminates between competing conformational clusters of ligand-bound protein structures. Using the determined X-ray crystal structure of PIM1 complexed to the compound 1-R as a control, we discuss the importance of including the protein flexibility inherent in the ensemble docking protocol, for the accuracy of the structure prediction of the bound state. A comparison of our ensemble docking results suggests that PIM1 and GSK-3β bind the two enantiomers in similar fashion, through two primary binding modes: conformation I, which is very similar to the conformation presented in the existing PIM1/compound 1-R crystal structure; conformation II, which represents a 180° flip about an axis through the NH group of the pyridocarbazole moiety, relative to conformation I. In contrast, the binding of the enantiomers to CDK2 is found to have a different structural profile including a suggested bound conformation, which lacks the conserved hydrogen bond between the kinase and the ligand (i.e., ATP, staurosporine, Ru-complex compound). The top scoring conformation of the inhibitor bound to CDK2 is not present among the top-scoring conformations of the inhibitor bound to either PIM1 or GSK-3β and vice-versa. Collectively, our results help provide atomic-level insights into inhibitor selectivity among the three kinases.
Figure
Top cluster of predicted conformations based on ensemble docking simulations of a Ruthenium based compound to protein kinases  相似文献   

12.
Three kinds of ATP analogues were synthesized. These ATP analogues can be classified into two conformations, i.e. syn and anti forms with respect to the N-glycosidic bond between adenine and ribose groups of ATP. 3'-O-(N-Methylanthraniloyl)-2-azidoadenosine 5'-triphosphate (MantN2(3)ATP) is recognized as the anti form, as ATP, and the other two, 3'-O-(N-methylanthraniloyl)-8-azidoadenosine 5'-triphosphate (MantN8(3)ATP) and 1,N6-etheno-8-azidoadenosine 5'-triphosphate (epsilon N8(3)ATP) are both syn forms. Mant and etheno groups are both fluorescent which allows detection of their binding to proteins. The photochemical binding of azido groups in ATP analogues to the myosin active site, examined in the presence and absence of ATP, showed that all the analogues bound to the site of myosin ATPase. These analogues also acted as substrates of the ATPase and were hydrolyzed in the active site, as judged by competitive inhibition of the ATPase and by their ATPase activities. Of these analogues, MantN2(3)ATP is very similar to ATP in divalent-cation dependence of its hydrolysis rate and in its ability to trap ADP in the active site with vanadate, while the other two are different from ATP in these respects. The photochemical binding sites of ATP analogues were localized by gel electrophoresis of trypsinized myosin ATPase with photocross-linked ATP analogues and/or by isolating the modified peptides. MantN2(3)ATP was found in the 23-kDa fragment which has a structure common to ATP-binding proteins, i.e. Gly-Xaa-Xaa-Gly-Xaa-Gly-Lys-Thr. Mant N8(3)ATP was found in a region of the 20-kDa fragment where actin is reported to attach.  相似文献   

13.
Despite their wide occurrence, proteoglycans (PGs) have never been isolated from the saliva of higher animals. We found that the Collocalia glycoproteins isolated from edible birds'-nests (the dried forms of regurgitated saliva of male Collocalia swiftlets) were rich in a PG containing nonsulfated chondroitin glycosaminoglycans (GAGs). We have devised a method to isolate a PG from the water extract of the white nest built by Aerodramus fuciphagus (white nest swiftlets) with a yield of 2-mg PG per gram nest. This PG contained 83% of carbohydrates, of which 79% were GalNAc and GlcUA (D-glucuronic acid) in an equimolar ratio. By using chondroitin AC lyase, the structure of GAGs in this PG was established to be chondroitin ( --> 4GlcUAbeta1 --> 3GalNAcbeta1 --> )(n) chains. The average molecular mass of the chondroitin chain was estimated to be 49 kDa by gel filtration. We have isolated a linkage region hexasaccharide, DeltaHexUAalpha1 --> 3GalNAcbeta1 --> 4GlcUAbeta1 --> 3Galbeta1 --> 3Galbeta1 --> 4Xyl, from this PG by chondroitinase ABC digestion to show that the GAGs in this PG are also linked to the core protein through the common tetrasaccharide linker, GlcUAbeta1 --> 3Galbeta1 --> 3Galbeta1 --> 4Xyl, found in various PGs. As water was not effective in extracting uronic acid-containing glycoconjugates from the black nest built by black nest swiftlets (A. maximus), we used 4 M guanidium chloride and anion-exchange chromatography in the presence of urea to extract and isolate about 30 mg of a chondroitin PG preparation from 10 g of the desialylated black nest. As the biological significance of chondroitin is still not well understood, bird's nest should become a convenient source for preparing this unique GAG to study its biological functions.  相似文献   

14.
The conformation of adenine nucleotides bound to bovine mitochondrial F1-ATPase was investigated using transfer nuclear Overhauser enhancement measurements. It is shown that all nucleotides investigated adopt a predominantly anti conformation when bound to the catalytic sites. Furthermore, the experiment suggests that 8-azido-ADP and 8-azido-ATP, which are predominantly in the syn conformation in solution, are in the anti conformation when bound to F1 catalytic sites.  相似文献   

15.
500 MHz H, homonuclear, intra-molecular, transferred Nuclear Overhauser Effect measurements have been performed on the bound forms of a classical opiate antagonist, nalorphine and an agonist, levorphanol at their respective binding sites in two different specific anti-opiate monoclonal antibody fragments. Based upon previous studies of opiate conformations in solution the results clearly show without extensive interpretation that one of these flexible haptens has the opposite (from solution) isomeric conformation in its bound form. For nalorphine the axial isomer of the N-allyl substituent is the bound form whereas in solution the equatorial isomer dominates at a ratio of 5:1. For levorphanol the bound form is that of equatorial N-methyl in accord with the low energy conformation in solution. In this preliminary report we discuss the initial measurements and results and their implications with respect to the conformations of flexible ligands at macromolecular binding sites including opiate receptors.  相似文献   

16.
Galactose oxidase (EC 1.1.3.9, GAO) was used to convert the C-6' OH of Galbeta(1 --> 4)Glcbeta-OBn (5) to the corresponding hydrated aldehyde (7). Chemical modification, through dehydratative coupling and reductive amination, gave rise to a small library of Galbeta(1 --> 4)Glcbeta-OBn analogues (9a-f, 10, 11). UDP-[6-(3)H]Gal studies indicated that alpha1,3-galactosyltransferase recognized the C-6' modified Galbeta(1 --> 4)Glcbeta-OBn analogues (9a-f, 10, 11). Preparative scale reactions ensued, utilizing a single enzyme UDP-Gal conversion as well as a dual enzymatic system (GalE and alpha1,3GalT), taking full advantage of the more economical UDP-Glc, giving rise to compounds 6, 15-22. Galalpha(1 --> 3)Galbeta(1 --> 4)Glcbeta-OBn trisaccharide (6) was produced on a large scale (2 g) and subjected to the same chemoenzymatic modification as stated above to produce C-6" modified derivatives (23-30). An ELISA bioassay was performed utilizing human anti-alphaGal antibodies to study the binding affinity of the derivatized epitopes (6, 15-30). Modifications made at the C-6' position did not alter the IgG antibody's ability to recognize the unnatural epitopes. Modifications made at the C-6" position resulted in significant or complete abrogation of recognition. The results indicate that the C-6' OH of the alphaGal trisaccharide epitope is not mandatory for antibody recognition.  相似文献   

17.
An agglutinin that has high affinity for GalNAcbeta1-->, was isolated from seeds of Wistaria sinensis by adsorption to immobilized mild acid-treated hog gastric mucin on Sepharose 4B matrix and elution with aqueous 0.2 M lactose. The binding property of this lectin was characterized by quantitative precipitin assay (QPA) and by inhibition of biotinylated lectin-glycan interaction. Of the 37 glycoforms tested by QPA, this agglutinin reacted best with a GalNAcbeta1-->4 containing glycoprotein (GP) [Tamm-Horsfall Sd(a+) GP]; a Galbeta1-->4GlcNAc containing GP (human blood group precursor glycoprotein from ovarian cyst fluid and asialo human alpha1-acid GP) and a GalNAcalpha1-->3GalNAc containing GP (asialo bird nest GP), but poorly or not at all with most sialic acid containing glycoproteins. Among the oligosaccharides tested, GalNAcalpha1-->3GalNAcbeta1-->3Galalpha1-->4Galbeta 1-->4Glc (Fp) was the most active ligand. It was as active as GalNAc and two to 11 times more active than Tn cluster mixtures, Galbeta1--> 3/4GlcNAc (I/II), GalNAcalpha1-->3(L-Fucalpha1-->2)Gal (Ah), Galbeta1-->4Glc (L), Galbeta1-->3GalNAc (T) and Galalpha1--> 3Galalpha-->methyl (B). Of the monosaccharides and their glycosides tested, p-nitrophenyl betaGalNAc was the best inhibitor; it was approximately 1.7 and 2.5 times more potent than its corresponding alpha anomer and GalNAc (or Fp), respectively. GalNAc was 53.3 times more active than Gal. From the present observations, it can be concluded that the Wistaria agglutinin (WSA) binds to the C-3, C-4 and C-6 positions of GalNAc and Gal residues; the N-acetyl group at C-2 enhances its binding dramatically. The combining site of WSA for GalNAc related ligands is most likely of a shallow type, able to recognize both alpha and beta anomers of GalNAc. Gal ligands must be Galbeta1-->3/4GlcNAc related, in which subterminal beta1-->3/4 GlcNAc contributes significantly to binding; hydrophobicity is important for binding of the beta anomer of Gal. The decreasing order of the affinity of WSA for mammalian structural carbohydrate units is Fp >/= multi-II > monomeric II >/= Tn, I and Ah >/= E and L > T > Gal.  相似文献   

18.
Sulfated forms of sialyl-Le(X) containing Gal-6-SO(4) or GlcNAc-6-SO(4) have been implicated as potential recognition determinants on high endothelial venule ligands for L-selectin. The optimal configuration of sulfate esters on the N-acetyllactosamine (Galbeta1-->4GlcNAc) core of sulfosialyl-Le(X), however, remains unsettled. Using a panel of sulfated lactose (Galbeta1-->4Glc) neoglycolipids as substrates in direct binding assays, we found that 6',6-disulfolactose was the preferred structure for L-selectin, although significant binding to 6'- and 6-sulfolactose was observed as well. Binding was EDTA-sensitive and blocked by L-selectin-specific monoclonal antibodies. Surprisingly, 6', 6-disulfolactose was poorly recognized by MECA-79, a carbohydrate- and sulfate-dependent monoclonal antibody that binds competitively to L-selectin ligands. Instead, MECA-79 bound preferentially to 6-sulfolactose. The difference in preferred substrates between L-selectin and MECA-79 may explain the variable activity of MECA-79 as an inhibitor of lymphocyte adhesion to high endothelial venules in lymphoid organs. Our results suggest that both Gal-6-SO(4) and GlcNAc-6-SO(4) may contribute to L-selectin recognition, either as components of sulfosialyl-Le(X) capping groups or in internal structures. By contrast, only GlcNAc-6-SO(4) appears to contribute to MECA-79 binding.  相似文献   

19.
Boxrud PD  Bock PE 《Biochemistry》2000,39(45):13974-13981
Binding of streptokinase (SK) to plasminogen (Pg) activates the zymogen conformationally and initiates its conversion into the fibrinolytic proteinase, plasmin (Pm). Equilibrium binding studies of SK interactions with a homologous series of catalytic site-labeled fluorescent Pg and Pm analogues were performed to resolve the contributions of lysine binding site interactions, associated changes between extended and compact conformations of Pg, and activation of the proteinase domain to the affinity for SK. SK bound to fluorescein-labeled [Glu]Pg(1) and [Lys]Pg(1) with dissociation constants of 624 +/- 112 and 38 +/- 5 nM, respectively, whereas labeled [Lys]Pm(1) bound with a 57000-fold tighter dissociation constant of 11 +/- 2 pM. Saturation of lysine binding sites with 6-aminohexanoic acid had no effect on SK binding to labeled [Glu]Pg(1), but weakened binding to labeled [Lys]Pg(1) and [Lys]Pm(1) 31- and 20-fold, respectively. At low Cl(-) concentrations, where [Glu]Pg assumes the extended conformation without occupation of lysine binding sites, a 23-fold increase in the affinity of SK for labeled [Glu]Pg(1) was observed, which was quantitatively accounted for by expression of new lysine binding site interactions. The results support the conclusion that the SK affinity for the fluorescent Pg and Pm analogues is enhanced 13-16-fold by conversion of labeled [Glu]Pg to the extended conformation of the [Lys]Pg derivative as a result of lysine binding site interactions, and is enhanced 3100-3500-fold further by the increased affinity of SK for the activated proteinase domain. The results imply that binding of SK to [Glu]Pg results in transition of [Glu]Pg to an extended conformation in an early event in the SK activation mechanism.  相似文献   

20.
Glutaraldehyde-polymerized human splenic galaptin, a beta-galactoside-binding lectin, was demonstrated to have enhanced hemagglutinating and asialofetuin binding activity relative to native dimeric galaptin when these lectins were present in solution. The polymerized lectin consisted primarily of 2-, 4- and 12-membered species after reductive alkylation. Both forms of galaptin bound, at 4 degrees C, to saturable B lymphoblastoid cell surface receptors. Estimates obtained by Scatchard analyses, with the binding data expressed in terms of 14.5 kDa subunit molarity, were 5 x 10(7) binding sites/cell with affinity constant Ka = 2.2 x 10(5) M for dimeric galaptin and 17 x 10(7) binding sites/cell with Ka = 3.4 x 10(5) M-1 for polymeric galaptin. Both forms of galaptin adsorbed to polystyrene with high efficiency; however, only plastic-adsorbed polymeric galaptin mediated adhesion of lymphoblastoid cells. Cell adhesion was inhibited by lactose. Plastic-adsorbed polymeric galaptin bound asialofetuin more efficiently than dimeric galaptin. Asialofetuin binding was inhibited 65% and 30-50% by lactose for plastic-adsorbed polymeric and dimeric galaptin, respectively. Native fetuin bound to the adsorbed dimeric galaptin in a lactose-insensitive manner. These data indicate that cell surface receptor-galaptin interaction is carbohydrate specific whereas polystyrene-adsorbed galaptin may demonstrate protein-protein interactions with soluble ligands.  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