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 15 毫秒
1.
We report dynamic light scattering measurements over a wide range of scattering vectors for fractionated samples of porcine submaxillary mucin (PSM) glycoproteins in two different solvents: 0.1M NaCl, and 6M GdnHCl. The relaxation spectrum has been successfully resolved into a slow mode corresponding to pure translational diffusion and a fast mode containing information on the relaxation times for intramolecular motion. Analysis of the slow mode permits a light scattering evaluation of the polydispersity of these high molecular weight mucin glycoprotein fractions. Determination of the longest intramolecular relaxation times tau 1 shows that these are much longer for the PSM fractions in 0.1M NaCl compared to 6M GdnHCl. These data are consistent with earlier studies showing that the chain conformation is the same in both solvents, but that in 0.1M NaCl, the PSM glycoprotein undergoes a self-association process that is end-to-end in nature. Since the tau 1 value is intimately related to the viscoelastic behavior of PSM solutions and gels, it is interesting to speculate that the end-to-end association process may be physiologically important.  相似文献   

2.
The glycoprotein mucin was isolated from the sputum of patients with chronic obstructive bronchitis. The fractionation procedure included treatment with 6M urea at pH 12.5 followed by gel filtration in 6M urea at neutral pH. (1) Using a low-shear Cartesian diver viscometer, we found that the mucin intrinsic viscosity equals (0.32 ± 0.03) L/g in 1000 mM NaCl solution increasing to (12 ± 3) L/g in 0.1 mM NaCl (pH 7 and 20°C). (2) The relaxation of electrically induced birefringence in mucin solutions was measured and the relaxation spectrum calculated using a Fourier-transform deconvoltion method. We found that the dominant relaxation time increased from 1 to 150 μs when the exitation pulse duration used was increased from 2 to 300 μs. (3) Mucin was vacuum-dried from glycerol-containing solutions followed by low-angle rotary shadowing and electron microscopy. Mucin was found to be unbranched, with contour lengths ranging from 300 to 2500 nm and with an average of 900 nm. Our result indicate that mucin is an extended and flexible molecule with Kuhn length 0.3–0.5% of the contour length.  相似文献   

3.
L M Soby  A M Jamieson  J Blackwell  N Jentoft 《Biopolymers》1990,29(10-11):1359-1366
The linear viscoelastic and rheological properties of high molecular weight ovine submaxillary mucin (OSM) solution have been investigated in terms of the Newtonian steady-flow viscosity [eta(gamma)], the complex oscillatory viscosity [eta*(omega)], and the storage and loss shear moduli [G'(omega) and G"(omega)]. It was observed that tau(gamma), eta*(omega), and G'(omega) are always higher when OSM is dissolved in 0.1M NaCl than when at the same concentration in 6M GdnHCl. This is consistent with previous observations that submaxillary mucins self-associate in 0.1M NaCl to form large aggregates, which are disrupted in 6M GdnHCl. As the OSM concentration increases, the appearance of a plateau shear modulus indicates the formation of a gel network in both solvents. The results suggest gelation involves specific intermolecular interactions, perhaps due to hydrophobic forces between interdigitated oligosaccharide side chains. The viscoelastic behavior of OSM solution at high concentration is thus similar to that reported in the literature for porcine gastric mucin (PGM). However, the OSM gels are mechanically weaker, having moduli that are an order of magnitude lower than those for PGM gels of comparable concentration. The oligosaccharide side chains of OSM consist of only 1-2 sugar units compared to 10-15 for PGM, but it appears that this is sufficient to allow for intermolecular interaction and the formation of weak gels.  相似文献   

4.
1. Binding of Ca-2+ to goblet cell mucin of rat small intestine was studied using equilibrium dialysis against 0.01 M Tris/HCl buffer (pH 7.4) and tracer amounts of 45-CaCl2. Binding was found to reach saturation at a Ca free -2+ concentration of 0.1--1.0 mM, to be independent of temperature (4-37 degrees C), and to increase with increasing pH (5.0-8.7). At low concentrations of Ca free -2+ (smaller than 0.03 mM) the binding curve was sigmoidal, suggesting positive cooperativity of binding sites and a possible change in the tertiary structure of the mucin. Binding was markedly reduced, and sigmoidicity abolished, by removal of sialic acid from the mucin, or by adding 0.14 M NaCl to the dialysis medium. This latter finding suggests that, in vivo, other cations would compete for Ca-2+ binding ligands. 2. Under conditions mimicking those used for binding studies, CaCl2 (10- minus 5 M) was found to cause a small increase (0.03 units) in the absorbance of mucin solutions, especially in the ultraviolet region, possibly indicating increased light scattering. No change in the solubility of the mucin was observed after the addition of CaCl2 (10- minus 6-10- minus 4 M). A significant decrease in viscosity of the mucin was noted, however, with the addition of CaCl2 (10- minus 6-10- minus 2 M). Together with the binding data, these findings suggested that during binding, Ca-2+ combines with negative charges on goblet cell mucin (especially those of sialic acid carboxyl groups) and induces contraction or folding of the macromolecule which promotes cooperative cation binding. No evidence was obtained to suggest that CaCl2 caused precipitation, polymerization or gelation of the mucin in 0.01 M Tris/HCl.?  相似文献   

5.
Proton magnetic longitudinal T1 relaxation times have been measured for acid (horse) ferrimyoglobin solutions [0.1 M NaCl and KH2PO4, 2 M NaCl and 1 M MgCl2] from 5°C to 35°C in dependence on myoglobin concentration up to 6 mM. The enhancement of the relaxation rate due to the paramagnetic haem iron. which is observed in this temperature range is compared with analogous data for the ferrihaemoglobin solution. The conclusion is that the protons exchanging from the haem pocket with bulk solvent are not those from the water molecule at the sixth ligand site of haem iron. The exchanging protons are more than 4 Å away from the haem iron being closer to it in ferrimyoglobin than in ferrihaemogiobin. This distance becomes larger in solutions with higher salt concentration, the largest difference between 0.1 M NaCl and 1 M MgCl2 being over one Angstrom unit. This indicates a conformational change of the haem pocket, possibly its tightening.  相似文献   

6.
Porcine submaxillary mucin (PSM) is a glycoprotein composed of a protein core and frequent, short oligosaccharide side chains. We report static and dynamic light scattering experiments and intrinsic viscosities for PSM in aqueous solvent systems. In 0.1M NaCl solution, the data suggest PSM exists as large, internally branched, highly hydrated, polydisperse aggregates that slowly dissociate to give a stable species of weight-average molecular weight (Mw) 7.4 × 106. In 6M GdnHCl solution, the noncovalent bonds between PSM molecules are broken, giving a highly elongated molecule of Mw = 2.0 × 106. The irreversible nature of this dissociation suggests that the forces that stabilize the native aggregates of PSM in 0.1M NaCl are specific in nature. On reduction of PSM with mercaptoethanol, the polydispersity decreases and Mw also decreases to 9 × 105. A discrete change is observed in the solution properties of PSM in 0.1M NaCl at a concentration of 2mg/mL, manifested by a sudden decrease in the translational diffusion coefficient, an increase in viscosity number, and a decrease in slope of the osmotic compressibility. We tentatively propose that a weak and reversible secondary association process occurs at this concentration, although a purely hydrodynamic interaction cannot be ruled out.  相似文献   

7.
1. The effects of morganic ions, electrolyte concentration, and pH on the appearance and volume of the isolated rat liver nucleus have been studied. Nuclei were isolated by differential centrifugation in a buffered salt-sucrose mixture at pH 7.1. Nuclear volumes were determined photographically. 2. In solutions of NaCl, of KCl, and in potassium phosphate buffers the nuclear volume decreased markedly with an increase in concentration from 0.001 M to 0.05 M but remained essentially constant with further increase in concentration to 1.0 M. The effects of CaCl(2) and MgCl(2) differed from those of NaCl and KCl in that a smaller volume was obtained in concentrations less than 0.15 M, and in the case of CaCl(2) an increase in volume was obtained in more concentrated solutions. The volume changes are considered to be due primarily to ionic effects on the nuclear colloids rather than to osmotic behavior. 3. Treatment of nuclei with DNAase prevented the characteristic volume changes resulting from ion effects, suggesting the importance of DNA in nuclear volume changes. 4. The optical changes in isolated nuclei in various concentrations of KCl, NaCl, CaCl(2), MgCl(2), and in potassium phosphate buffers as observed under phase contrast illumination are described. CaCl(2) gave the most marked nuclear changes from the conditions in the uninjured cell and caused shrinkage and granulation in 0.001 M concentration. The effects of CaCl(2) were also manifested in 0.88 M sucrose, in mixtures with monovalent salts, and in serum. Changes in nuclear volume and optical appearance which occurred in salt solutions and in 0.1 N HCl were readily reversible. 5. Nuclear volume remained constant between pH 8.91 and 5.12 and decreased in more acid solutions. 6. Sucrose had no appreciable osmotic effect, and in hyperosmotic solution. (0.88 M) nuclei showed swelling and rupture comparable to that in distilled water. 7. The results are considered in relation to the requirements of nuclear isolation media. 8. Rat liver nuclei isolated in a buffered salt-sucrose medium by differential centrifugation exhibited a pattern of size distribution similar to that of fixed nuclei but were of considerably larger volume. The ratio of the volumes of the peak frequencies of the two chief size groups was 1:1.9.  相似文献   

8.
J R Lyerla  D A Torchia 《Biochemistry》1975,14(23):5175-5183
13C relaxation parameters, T1, line width, and NOE, have been determined for backbone carbons of ligamentum nuchae elastin swollen by 0.15 M NaCl, 0.15 M NaCl-formamide, 0.15 M NaCl-ethanol, dimethyl sulfoxide, and formamide. The data have been analyzed in terms of (a) a single correlation time model and (b) a model employing a log-chi2 distribution of correlation times used by Schaefer (1973) to analyze solid cis-polyisoprene 13C relaxation data. Employing the latter mode, one obtains an approximately self-consistent quantitative analysis of all the elastin data. An average backbone correlation time, tau, of ca. 2 nsec is calculated for elastin swollen in the presence of polar organic solvents at 37 degrees, in approximate agreement with tau of 0.4 nsec obtained for bulk cis-polyisoprene at 35 degrees. The influence of solvent and temperature on elastin spectra indicate that the larger tau value (approximately 80 nsec) obtained for elastin swollen by 0.15 M NaCl at 37 degrees is a consequence of weak interchain polar and hydrophobic interactions, a result which is in accord with the reported viscoelastic behavior exhibited by water-swollen elastin at 37 degrees. The results obtained further suggest that Gly, Pro, and Val residues are significantly more mobile than Ala residues, which are located in the cross-link regions. Hence, the NMR data support the view that water-swollen elastin is composed of a network of mobile chains, except possibly in the cross-link regions.  相似文献   

9.
The steady-state gating of individual batrachotoxin-modified sodium channels in neutral phospholipid bilayers exhibits spontaneous, reversible changes in channel activation, such that the midpoint potential (Va) for the gating curves may change, by 30 mV or more, with or without a change in the apparent gating valence (za). Consequently, estimates for Va and, in particular, za from ensemble-averaged gating curves differ from the average values for Va and za from single-channel gating curves. In addition to these spontaneous variations, the average Va shifts systematically as a function of [NaCl] (being -109, -88, and -75 mV at 0.1, 0.5, and 1.0 M NaCl), with no systematic variation in the average za (approximately 3.7). The [NaCl]-dependent shifts in Va were interpreted in terms of screening of fixed charges near the channels' gating machinery. Estimates for the extracellular and intracellular apparent charge densities (sigma e = -0.7 and sigma i = -0.08 e/nm2) were obtained from experiments in symmetrical and asymmetrical NaCl solutions using the Gouy-Chapman theory. In 0.1 M NaCl the extracellular and intracellular surface potentials are estimated to be -94 and -17 mV, respectively. The intrinsic midpoint potential, corrected for the surface potentials, is thus about -30 mV, and the standard free energy of activation is approximately -12 kJ/mol. In symmetrical 0.1 M NaCl, addition of 0.005 M Ba2+ to the extracellular solution produced a 17-mV depolarizing shift in Va and a slight reduction in za. The shift is consistent with predictions using the Gouy-Chapman theory and the above estimate for sigma e. Subsequent addition of 0.005 M Ba2+ to the intracellular solution produced a approximately 5-mV hyperpolarizing shift in the ensemble-averaged gating curve and reduced za by approximately 1. This Ba(2+)-induced shift is threefold larger than predicted, which together with the reduction in za implies that Ba2+ may bind at the intracellular channel surface.  相似文献   

10.
Conformation and dynamics of short DNA duplexes: (dC-dG)3 and (dC-dG)4   总被引:2,自引:0,他引:2  
Natural abundance 13C NMR spectra of duplexed (dC-dG)3 and (dC-dG)4 exhibit resolved resonances for most of the carbons at 0.1M NaCl in aqueous solution. Large transitions in chemical shift for many of the hexamer carbons (up to 1.8 ppm) are observed in variable temperature measurements. Determination of spin-lattice relaxation times and nuclear Overhauser enhancements in 0.1M NaCl indicate that the duplexes tumble almost isotropically, with overall correlation times near 5 nsec; the sugar carbons experience more rapid local motions than do the base carbons. The relaxation data are also consistent with the most rapid local motions occurring at the chain-terminal residues, especially in the Cyd(1) sugar. 4M NaCl causes changes in the 13C chemical shifts of most of the guanine base carbons, and rearrangements in the deoxyribose carbon shifts; this is consistent with changes predicted by a salt-induced B to Z transition, viz. conversion of the guanylates from the anti to syn range about the glycosyl bond, and from the S to N pseudorotational state of the deoxyribose ring.  相似文献   

11.
The study of water relaxation rates of solutions of melanins with paramagnetic metal ions provides a powerful tool for investigating the binding sites and for obtaining useful structural and dynamic information. The measure of 1H- and 2H-longitudinal and transverse relaxation rates at a single, high-magnetic field for H2O/D2O (80:20) solutions allows the determination of tau c, tau R, tau e, tau M, r, q, and Z (the outer sphere contribution to the overall relaxation rate) for Mn(II)-L-Dopa melanin system.  相似文献   

12.
The thermal depolymerization of porcine submaxillary mucin   总被引:2,自引:0,他引:2  
The time dependence of the molecular weight, radius of gyration, and hydrodynamic size distribution for porcine submaxillary mucin (PSM) in solution have been studied using static and dynamic light scattering. The weight average molecular weight (Mw) of PSM in 6 M guanidine HCl, pH 7, is initially 3 X 10(6) and decreases with time in three phases: rapidly from 3-2 X 10(6), less rapidly from 2-0.9 X 10(6), and slowly below 0.9 X 10(6). The rates of decrease are much greater at pH 2. The energy of activation associated with each phase is 20 kcal/mol, which is similar to that reported for peptide bond cleavage at an aspartic acid residue. Addition of mercaptoethanol to PSM in 6 M guanidine HCl leads to a rapid decrease in Mw to 0.9 X 10(6), followed by a very slow further decrease. These results suggest that native PSM consists of subunits (Mw = 0.9 X 10(6] that are linked by disulfide bonds to form dimers (Mw = 2 X 10(6] and then higher aggregates. This cross-linking appears to occur at unglycosylated regions of the protein core, which are believed to be richer in aspartic acid than the rest of the molecule.  相似文献   

13.
A comparison of 17O and 2H NMR relaxation rates of water in lysozyme solutions as a function of concentration, pH/pD, and magnetic field suggests that only 17O monitors directly the hydration of lysozyme in solution. NMR measurements are for the first time extended to 11.75 T. Lysozyme hydration data are analyzed in terms of an anisotropic, dual-motion model with fast exchange of water between the "bound" and "free" states. The analysis yields 180 mol "bound" water/mol lysozyme and two correlation times of 7.4 ns ("slow") and 29 ps ("fast") for the bound water population at 27 degrees C and pH 5.1, in the absence of salt, assuming anisotropic motions of water with an order parameter value for bound water of 0.12. Under these conditions, the value of the slow correlation time of bound water (7.4 ns) is consistent with the value of 8 ns obtained by frequency-domain fluorescence techniques for the correlation time associated with the lysozyme tumbling motion in solutions without salt. In the presence of 0.1 M NaCl the hydration number increases to 290 mol/mol lysozyme at pD 4.5 and 21 degrees C. The associated correlation times at 21 degrees C in the presence of 0.1 M NaCl are 4.7 ns and 15.5 ps, respectively. The value of the slow correlation time of 4.7 ns is consistent with the calculated value (4.9 ns) for the lysozyme monomer tumbling in solution. The systematic deviations of the relaxation rates, estimated with the single-exponential approximation, from the theoretical, multiexponential nuclear (I' + 1/2) spin relaxation are evaluated at various frequencies for 17O (I = 5/2) with the first-order, linear approximation (25). All NMR relaxation data for hydrated lysozymes are affected by protein activity and are sensitive both to the ionization of protein side chains and to the state of protein aggregation.  相似文献   

14.
Conformation of mucous glycoproteins in aqueous solvents   总被引:5,自引:0,他引:5  
Light-scattering techniques have been used to measure the z-average radius of gyration Rg z-average translational diffusion coefficient Dt and weight–average molecular weight Mw of porcine submaxillary mucin (PSM) in solution. PSM isolated at low shear in the presence of protease inhibitors has a Mw about twice as large as a sample prepared without these precautions. The former sample has a Mw of 17 × 106 in 0.1M NaCl, which decreases to 8 × 106 in 6M guanidine hydrochloride (GdnHCl) and then to 2 × 106 on addition of 0.1M mercaptoethanol to the 6M GdnHCl solution. The Rg or D values obtained for PSM in this work superimpose with those of other authors for different mucin glycoproteins, leading to linear log–log relationships to the molecular weight of the protein core. Comparison of these results with those in the literature for denatured proteins suggest that mucins are linear random coils in which the protein core is stiffened by the presence of the oligosaccharide side chains. The length of the oligosaccharides and the nature of the solvent have little effect on the extension of the protein core. This suggests that the stiffness of the protein core is maintained by steric repulsion of the residues at the beginning of the oligosaccharide chains.  相似文献   

15.
The amounts of sugar solution consumed by Drosophila melanogaster flies were determined. Starved and desiccated flies of a wild type strain (QA) consume 7?9 × 10?2 λ of a 0.3 M sucrose solution per fly during the first hour and less later. They consume more of the 0.3 M sucrose solution than of the more diluted and the more concentrated solutions. In preference-aversion tests the flies discriminated between water and various sugar solutions, and between different sugar concentrations. Contrary to other fly species these flies did not prefer 0.05 M fructose over 0.05 M glucose. 0.3–0.5 M NaCl added to 0.1 M sucrose turned a preference over 0.01 sucrose into an aversion. A mutant, Lot-94, selected for its increased consumption of a 1 M NaCl solution was found to consume more of all test solutions. The amount of NaCl that had to be added to 0.1 M sucrose to turn the preference over 0.01 M sucrose by the mutant flies into aversion was not different from that found for the wild type flies.  相似文献   

16.
Nejad  Mohssen S. 《Chemical senses》1986,11(3):283-293
A comparison of the integrated responses of the rat's greatersuperficial petrosal (GSP) and chorda tympani (CT) nerves toa number of taste stimuli was studied. The GSP nerve of therat was very responsive to the chemical stimulation of the oralcavity. Among the selected stimuli related to the four basictaste qualities, 0.5 M sucrose produced the greatest neuralresponse in the GSP nerve, whereas, 0.1 M NaCl produced thegreatest in the CT nerve. The GSP nerve integrated responseto 0.5 M sucrose solution was approximately three times as greatin magnitude as that to a 0.1 M NaCl solution. The neural responsemagnitude of the GSP and CT nerves were as follows: GSP nerve;0.5 M sucrose >0.02 M Na-saccharin >0.05 M citric acid>0.1 M NaCl > 0.01 M quinine-HCl. CT nerve; 0.1 M NaCl> 0.05 M citric acid > 0.02 M Na-saccharin > 0.01 Mquinine-HCl >0.5 M sucrose. The response magnitudes of theGSP nerve to 0.3 M chloride salt solutions were: LiCl > CaCl2> NaCl > NH4Cl > KCl, whereas the response magnitudesof the CT nerve to the above salts were: LiCl > NaCl >NH4Cl > CaCl2 > KCl. All 0.5 M solutions of the selectedsugars (sucrose, rhamnose, galactose, lactose, fructose, -methyl-D-glucoside,xylose, mannose, arabinose, maltose, sorbose and glucose) evokedneural responses in both GSP and CT nerves. The order of theresponse magnitudes of the GSP nerve to the selected sugarswas similar to that of the CT nerve but the absolute magnitudesof the GSP nerve were greater.  相似文献   

17.
Effect of Na+, K+ and Ca2+ on gel transition temperature (Tg) of gelatin hydrogels (5%, w/v) has been studied by oscillatory rheology in the salt concentration range I = 0.01-0.1 M, which showed increase in Tg with salt concentration with the trend for Tg showing Ca2+ > K+ > Na+. The dynamic light scattering (DLS) measurements in the sol state (T>Tg) showed two distinct relaxation modes whereas only a gel mode was observed in the gel state in all the samples which contained significant amount of heterodyne contribution. Low frequency (1.5 rad/s) isochronal storage modulus data revealed the formation of strong gel in presence of CaCl2 compared to that of NaCl and KCl situations. The slow mode relaxation and heterodyne parameter obtained from DLS data indicate the presence of larger clusters in Ca2+ gels.  相似文献   

18.
Aqueous solutions of sodium caseinate and xanthan at pH 7 and containing 0.1 M NaCl, and their mixtures were investigated using dynamic light scattering. Sodium caseinate solutions showed a bimodal distribution of relaxation rates; with the aggregate peak distribution predominating. Xanthan solutions showed a single distribution at low concentrations (≤0.06 wt.%) and a bimodal distribution at higher concentrations. The sodium caseinate–xanthan mixture modes were independent of the total biopolymer concentration, and behaved as a superposition of sodium caseinate solution alone and xanthan solution alone. This indicates that there is no interaction between xanthan and sodium caseinate in the range of concentrations considered in this study.  相似文献   

19.
《Biorheology》1995,32(4):431-446
Rheological methods have been used to investigate the intermolecular interactions of porcine submaxillary mucins (PSM) in solution. PSM is a high molecular weight glycoprotein consisting of a linear, semi-flexible protein backbone to which a large number of oligosaccharides (1–5 saccharide units) are attached as side chains. Concentrated aqueous solutions of PSM containing different amounts of guanidine hydrochloride (GdnHCl) were subjected to both controlled stress and controlled strain rheological analyses. In the absence of GdnHCl, PSM solutions exhibit viscoelastic properties characteristic of a gel: the storage modulus, G′, is much larger than the loss modulus, G″, at all deformation frequencies, and the compliance is 100% recoverable at small stresses, indicative of strong intermolecular interactions. In 3.0 M aqueous GdnHCl, PSM forms a viscoelastic solution, with G″ > G′ at all frequencies and a relatively small recoverable compliance, pointing to disruption of the intermolecular interactions by the chaotropic salt. Intermediate behavior is observed in 1.5 M GdnHCl, characteristic of a marginal gel: G′ ≈ G″ and greater than 50% recoverable compliance. In dilute solution, PSM behaves viscoelastically as a typical polyelectrolyte. However, concentrated solutions are turbid, the turbidity decreasing as GdnHCl is added, indicating that extensive intermolecular association accompanies the gelation process. The results show that although PSM is secreted in nature as a viscous solution, it can form gels that are similar to those of tracheobronchial and gastric mucins, and suggest common features to the gelation mechanism, with the strength of the gel correlated with the length of the oligosaccharide side chains.  相似文献   

20.
T A Gerken  N Jentoft 《Biochemistry》1987,26(15):4689-4699
Nearly all of the resonances in the 13C NMR spectrum of porcine submaxillary mucin glycoprotein (PSM) have been assigned to the peptide core carbons and to the carbons in the eight different oligosaccharide side chains that arise from the incomplete biosynthesis of the sialylated A blood group pentasaccharide (alpha-GalNAc(1-3) [alpha-Fuc(1-2)]-beta-Gal(1-3) [alpha-NeuNGl(2-6)]- alpha-GalNAc-O-Ser/Thr). By use of these assignments, a nearly complete structural analysis of intact PSM has been performed without resorting to degradative chemical methods. Considerable structural variability in the carbohydrate side chains was observed between mucins obtained from different animals, while no variability was observed between glands in a single animal. The dynamics of the PSM core and carbohydrate side chains were examined by using the carbon-13 nuclear magnetic resonance relaxation times and nuclear Overhauser enhancements of each assigned carbon resonance. The peptide core of PSM exhibits internal segmental flexibility that is virtually identical with that of ovine submaxillary mucin (OSM), whose carbohydrate side chain consists of the alpha-NeuNAc(2-6)alpha-GalNAc disaccharide. The longer oligosaccharide side chains of PSM, therefore, have no significant effect on peptide core mobility compared to the shorter side chains of native OSM or asialo-OSM. Although the dynamics of the shorter carbohydrate side chains shared by both OSM and PSM appear to be identical, the A and H blood group structures in PSM have reduced mobilities, indicating that the glycosidic linkages of the terminal sugars in these determinants are relatively inflexible. These results differ from most reports of glycoprotein dynamics, which typically find the terminal carbohydrate residues to be undergoing rapid internal rotation about their terminal glycosidic bonds. The results reported here are consistent with previous studies on the conformations of the A and H determinants derived from model oligosaccharides and further indicate that the conformations of these determinants are unchanged when covalently bound to the mucin peptide core. In spite of their carbohydrate side-chain heterogeneity, mucins appear to be ideal glycoproteins for the study of O-linked oligosaccharide conformation and dynamics and for the study of the effects of glycosylation on polypeptide conformation and dynamics.  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