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.
Effects of sodium chloride (NaCl), guanidine hydrochloride (GuHCl), or sucrose on the viscoelasticity of sodium hyaluronate (NaHA) solutions were studied. NaCl and GuHCl decreased both storage and loss moduli, while sucrose increased both moduli. The critical concentration C* was determined as an inflection point in the plot of zero shear specific viscosity vs concentration for NaHA solutions with and without NaCl, GuHCl, or sucrose. It is suggested that sodium ions or guanidinium ions shield the electrostatic repulsion of NaHA molecules, hence reduce the coil dimension, and C* shifted to higher concentrations. However, sucrose enhances the entanglement coupling between NaHA molecules and retards the disentanglement of molecular chains or promotes to create hydrogen bonds, and then C* for NaHA solutions with sucrose shifts to lower concentrations. This is in agreement with the results of light scattering measurements in the presence of 0.2M NaCl. Both the radius of gyration and hydrodynamic radius of NaHA were reduced in dilute solutions by the addition of sucrose, and added sucrose enhances the interaction between NaHA monomer units. In the case of concentrated NaHA solution, such interactions result to increase the storage and loss moduli because of the enhancement of temporary network formation. © 1999 John Wiley & Sons, Inc. Biopoly 50: 23–34, 1999  相似文献   

2.
Mass transfer coefficients and interfacial areas were determined for the aeration of aqueous solutions in a turbine agitated vessel. The mass transfer coefficients measured for water without additive and for sodium chloride solutions matched very well to measurements in the literature for air bubbles of the same diameter in free rise. Thus the only effect of agitation was to determine the bubble size which then in turn set the coefficient. Two surface active agents were studied: sodium dodecyl sulfate and Dow Corning Antifoam C. The rate of mass transfer increased with the former additive but decreased with the latter; however, the mass transfer coefficient was the exact same function of bubble diameter in both cases and the different rates are attributed to the quite different effects on interfacial area.  相似文献   

3.
The vesicle-to-micelle transition of egg phosphatidylcholine LUVs induced by octylglucoside was studied in buffers with 0-4 M sodium chloride, sucrose or urea. We used both light scattering and fluorescent probes to follow the lipid-detergent complexes in these buffers. The vesicle-to-micelle transition process was fundamentally the same in each solute. However, the detergent-to-lipid ratio required for micelle formation shifted in ways that depended on the aqueous solute. The partitioning of octylglucoside between the vesicles and the aqueous phase was primarily determined by the change in its critical micelle concentration (cmc) induced by each solute. Specifically, the cmc decreased in high salt and sucrose buffers but increased in high concentrations of urea. Cmc for two additional nonionic detergents, decyl- and dodecyl-maltoside, and three zwittergents (3-12, 3-14 and 3-16) were determined as a function of concentration for each of the solutes. In all cases NaCl and sucrose decreased the solubility of the detergents, whereas urea increased their solubilities. The effects clearly depended on acyl chain length in urea-containing solutions, but this dependence was less clear with increasing NaCl and sucrose concentrations. The contributions of these solutes to solubility and to interfacial interactions in the bilayers, pure and mixed micelles are considered.  相似文献   

4.
Various molecular parameters, which characterize sodium hyaluronate in 0.2M NaCl solution, were obtained at 25°C by means of the static and dynamic light scattering and low shear viscometry over the molecular weight range of 5.94–627 × 104. Molecular weight distribution was obtained by using the Laplace inversion method of the autocorrelation function of the scattered light intensity and by Yamakawa theory for the wormlike chain with the stiff chain parameters for sodium hyaluronate in 0.2M NaCl (persistence length, chain diameter, molar mass per unit contour length, and the excluded‐volume strength). The molecular weight distribution thus obtained reproduced the solution properties of sodium hyaluronate well. Especially, the intrinsic viscosity showed a good agreement over four orders of molecular weight with Yamakawa theory combined with the Barrett function. Sodium hyaluronate in 0.2M NaCl solution is well expressed by the wormlike chain model affected by the excluded‐volume effect with the persistence length of 4.2 nm. © 1999 John Wiley & Sons, Inc. Biopoly 50: 87–98, 1999  相似文献   

5.
The swelling and viscoelastic properties of purified elastin were studied in aqueous solutions of superswelling agents or osmotic deswelling agents to develop models to study the behavior of elastin at frequencies not easily accessible by direct measurement. Increasing the concentration of any of the deswelling solutes (glucose, sucrose, sodium chloride, ammonium sulphate, dextran, and polyethylene glycol) increased the tensile storage and loss moduli. The viscoelastic behavior was independent of solute when compared on the basis of swelling behavior. The data collected at various solute concentrations at 37°C could be reduced to one master curve, and the master curves for elastin in each of the deswelling solutes were themselves superposable. The ability to reduce the data indicates that dehydration can be used to model elastin's viscoelastic behavior at high frequencies or over short times. The viscoelastic behavior of elastin in the superswelling agents [potassium thiocyanate (KSCN), dimethyl sulfoxide (DMSO), and ethylene glycol (EG)] depended on the solute and was independent of swelling behavior. In KSCN the behavior of elastin seemed to be a continuation of the pattern established by the deswelling agents in that an increase in swelling was accompanied by a decrease in both moduli, and the viscoelastic spectra were reducible to one master curve. In high concentrations of DMSO and EG the spectra were not reducible. KSCN appears a suitable superswelling solute to model elastin's viscoelastic behavior at low frequencies or over long times. © 1996 John Wiley & Sons, Inc.  相似文献   

6.
Methods of potentiometry, turbidimetry, colorimetry, IR spectroscopy, and element analysis were used to investigate the conditions of formation and the properties of non-stoichiometric polyelectrolyte complexes of chitosan hydrochloride (CHC) and sodium dextransulfate (the molecules of both polysaccharides appear as semirigid chains, but their charges are opposite). It was determined that the complexes' formation of polyelectrolytes studied is predominantly electrostatic in the presence of urea. As was also found turbidity and stability of the polycomplexes solutions depended markedly on pH value of CHC and a nature of the low-molecular-weight salts added. The complexes obtained were soluble in water, aqueous urea, and water-organic mixtures. The extent of solubility depended on the composition of the complexes and could be influenced by addition of appropriate concentrations of certain low-molecular-weight salts.  相似文献   

7.
The hydrolytic products of a chitinase purified from Nocardia orientalis were examined on reduced (GIcNAc)n(n = 2~6). The rate of hydrolysis on reduced (GlcNAc)4^6 increased with increasing chain-length of A-acetylglucosamine residues, but the enzyme did not act on reduced (G1cNAc)2 or reduced (GlcNAc)3. Based on the analysis of the frequency distribution of bond cleavage on PNP-(GIcNAc)?(n = 2 ~ 5) in the initial hydrolysis, the enzyme was shown to release predominantly (G1cNAc)2 from the nonreducing end of each substrate. The enzyme, which is essentially a hydrolase, also catalyzed a transglycosylation reaction in an excess of (GlcNAc)4 as an initial substrate. In this case, the addition of ammonium sulfate to the reaction system resulted in a significant increase in (G1cNAc)6 production. The yield of the hexasaccharide was about 34% of the chitinase-catalyzed net decrease of (GlcNAc)4. The rate of the transglycosylation in the presence of ammonium sulfate greatly depended on the salt concentration, the temperature, and the substrate concentration.  相似文献   

8.
The influence of added salts on the dynamic viscoelastic properties are investigated for aqueous solutions of alginates that have various molecular weights and mannuronate/guluronate (M/G) ratios. The dynamic moduli of the systems increase with increasing concentration of the added salt in the low-frequency region. The effect is notable in the order of KCl < NaCl < MgCl2 ? CaCl2. The values of the dynamic moduli in the rubbery plateau are independent of the addition of the salts, irrespective of the M/G ratio of the alginate. These facts strongly suggest that the structure that is formed by the interaction between the alginates and the metal ions does not work as cross-linking points but as heterogeneous relaxation units having a relatively long relaxation time from a rheological viewpoint.  相似文献   

9.
The effect of surfactants on the "fluidity" of dipalmitoylphosphatidylcholine (DPPC) vesicle membrane was studied by means of the fluorescence depolarization technique with fatty acid fluorescent probes, in which the anthroyloxy group is introduced at different positions along the acyl chain. Three types of surfactants were examined; anionic sodium alkylsulfates, cationic alkyltrimethylammonium chlorides, and non-ionic alkanoyl-N-methylglucamides (MEGA-n). Perturbing effects of the surfactants depended on both the alkyl chain-length and the type of head group. Sodium alkylsulfates with octyl- and decyl-chain and alkyltrimethylammonium chlorides with octyl-, decyl- and dodecyl-chain did not affect the membrane fluidity when incorporated in the membrane, whereas sodium dodecylsulfate and tetradecyltrimethylammonium chloride decreased the membrane fluidity at both gel and liquid crystalline states of the membrane. All the MEGA series surfactants decreased the membrane fluidity, whose perturbing potency was in the order of MEGA-8 less than MEGA-9 approximately equal to MEGA-10. The perturbation at different depths in the membrane by sodium dodecylsulfate and MEGA-9 was also examined. No significant change in the fluidity gradient across the membrane was induced by the addition of these surfactants.  相似文献   

10.
The conformation of the mucopolysaccharides. Hyaluronates   总被引:9,自引:7,他引:2       下载免费PDF全文
X-ray-diffraction patterns of hyaluronate fibres from a variety of sources were obtained. Sodium hyaluronate gives well-defined patterns which index on a hexagonal unit cell with dimensions a=1.17+/-nm and a fibre repeat-distance of 2.85+/-0.03nm. A further form of sodium hyaluronate is produced by annealing at 60 degrees C in 75% relative humidity. This stable state indexes on a hexagonal unit cell of unchanged fibre repeat-distance but with a=1.87nm. The chain conformation is a threefold helix. Analysis of these diffraction patterns led to two tentative structures for sodium hyaluronate, involving different packing of the polysaccharide chains. The significance of side-chain interaction is discussed. Hyaluronic acid produces an X-ray pattern different from that obtained with the sodium salt. The fibre repeat-distance is 1.96+/-0.02nm and the unit cell appears to be monoclinic. The chain conformation is a twofold helix and conformational change between free acid and monovalent salt is discussed. These findings, together with model-building experiments, are interpreted as indicating a highly ordered structure, and the physical properties of hyaluronate solutions with regard to molecular shape and polyelectrolyte behaviour are rationalized.  相似文献   

11.
Wang J  Yan Z  Zhuo K  Lu J 《Biophysical chemistry》1999,80(3):179-188
The apparent molar volumes V(2,phi) have been determined for glycine, DL-alpha-alanine, DL-alpha-amino-n-butyric acid, DL-valine and DL-leucine in aqueous solutions of 0.5, 1.0, 1.5 and 2.0 mol kg(-1) sodium acetate by density measurements at 308.15 K. These data have been used to derive the infinite dilution apparent molar volumes V(0)(2,phi) for the amino acids in aqueous sodium acetate solutions and the standard volumes of transfer, Delta(t)V(0), of the amino acids from water to aqueous sodium acetate solutions. It has been observed that both V(0)(2,phi) and Delta(t)V(0) vary linearly with increasing number of carbon atoms in the alkyl chain of the amino acids. These linear correlations have been utilized to estimate the contributions of the charged end groups (NH(3)(+), COO(-)), CH(2) group and other alkyl chains of the amino acids to V(0)(2,phi) and Delta(t)V(0). The results show that V(0)(2,phi) values for (NH(3)(+), COO(-)) groups increase with sodium acetate concentration, and those for CH(2) are almost constant over the studied sodium acetate concentration range. The transfer volume increases and the hydration number of the amino acids decreases with increasing electrolyte concentrations. These facts indicate that strong interactions occur between the ions of sodium acetate and the charged centers of the amino acids. The volumetric interaction parameters of the amino acids with sodium acetate were calculated in water. The pair interaction parameters are found to be positive and decreased with increasing alkyl chain length of the amino acids, suggesting that sodium acetate has a stronger dehydration effect on amino acids which have longer hydrophobic alkyl chains. These phenomena are discussed by means of the co-sphere overlap model.  相似文献   

12.
Synovial fluid is a approximately 0.15% (w/v) aqueous solution of hyaluronic acid (HA), a polysaccharide consisting of alternating units of GlcA and GlcNAc. In synovial fluid of patients suffering from rheumatoid arthritis, HA is thought to be degraded either by radicals generated by Fenton chemistry (Fe2+/H2O2) or by NaOCl generated by myeloperoxidase. We investigated the course of model reactions of these two reactants in physiological buffer with HA, and with the corresponding monomers GlcA and GlcNAc. meso-Tartaric acid, arabinuronic acid, arabinaric acid and glucaric acid were identified by GC-MS as oxidation products of glucuronic acid. When GlcNAc was oxidised, erythronic acid, arabinonic acid, 2-acetamido-2-deoxy-gluconic acid, glyceric acid, erythrose and arabinose were formed. NaOCl oxidation of HA yielded meso-tartaric acid; in addition, arabinaric acid and glucaric acid were obtained by oxidation with Fe2+/H2O2. These results indicate that oxidative degradation of HA proceeds primarily at glucuronic acid residues. meso-Tartaric acid may be a useful biomarker of hyaluronate oxidation since it is produced by both NaOCl and Fenton chemistry.  相似文献   

13.
The CD spectrum of an enzymatically derived sodium hyaluronate (NaHA) segment preparation with chain length 18 ± 3 disaccharide units [NaHAseg, ( NaGlcUA GlcNAc)15–20°. NaGlcUA, sodium D -glucuronate; GlcNAc, 2-acetamido-2-deoxy-D -glucose] in H2O was recorded to 180 nm using a computer-controlled vacuum-uv CD instrument. Near 190 nm the spectrum is of low intensity, similar to the sum of the free monosaccharide contributios, attributed to the π–π* transitions of the acetamido and carboxylate substituents. In contrast, much smaller oligosaccharides, also derived from high-molecular-weight NaHA by enzymatic digestions, show CD spectra in H2O with prominent bands centered near 190 nm. The oligosaccharide spectra can be matched as linear combinations of interior sugar residue (? NaHAseg) and end sugar residue CD contributions. End residues from oligosaccharides of the type (NaGlcUA-GlcNAc)n show a negative CD band near 190 nm. End residues from oligosaccharides of the reverse sequence (GlcNAc-NaGlcUA)n show a positive CD band near 190 nm. Averaging of the two end-residue spectral contributions yields an approximate match for the spectrum of NAHAseg below 200 nm. It is proposed that the low intensity CD of NaHA in the π–π* region is the result of large-magnitude, oppositely signed contributions, which can be visulized by studying oligosaccharides.  相似文献   

14.
Rheological properties of concentrated chitosan aqueous solutions and gels in the presence of different organic and inorganic acids were investigated. Viscosities of the solutions increased with polymer concentration and degree of ionization. Strong gels were obtained at pH around 2 with oxalic, phosphoric and sulfuric acids. Gelation was favored by simple and short chain length acids and was governed by ionic interactions. The gels could be distinguished from solutions by the frequency independence of their dynamic moduli and their high apparent activation energy for flow.  相似文献   

15.
Glycoprotein-glycans have recently been implicated to play a variety of functional roles. The same glycan chain have been found complexed with proteins of diverse functions. In this article two such glycan chains found attached to Fc regions of immunoglobulin G and immunoglobulin M have been studied. An extensive simulated annealing procedure have been adopted to arrive at a low-energy minimum of the two oligosaccharides. Molecular dynamics simulations have been performed to study the flexibility of the glycosidic linkages. It was found that both glycan chains can undergo conformational transitions and adopt folded and extended conformations. The two β(1–2) linkages of complex-type glycan had been found to prefer different conformational regime and the terminal fucose linked to the GlcNAc residue drastically modifies the GlcNAc β(1–4)GlcNAc linkage conformation. In the high-mannose type glycan chain α(1–3) linkages can induce flexibility in addition to the α(1–6) linkages. The results have been compared with recent experimental nmr and fluorescence energy transfer data. © 1998 John Wiley & Sons, Inc. Biopoly 45: 177–190, 1998  相似文献   

16.
The critical micelle concentration (CMC) of four synthetic phosphatidylcholines (containing two hexanoyl, heptanoyl, octanoyl or nonanoyl residues respectively) in aqueous solutions have been determined by surface tension measurements. The dependence of the CMC on the chain length is discussed on the basis of the mass action model for micelle formation. For the three higher homologues a contribution of 1.08 kT per CH2 group to the standard free energy of micellisation is found. The change in this free energy in going from the dihexanoyl- to the diheptanoyllecithin is somewhat larger (1.2 kT per CH2 group).The influence of high concentrations (several moles per liter) of simple electrolytes on the CMC is interpreted as a salting-out of nonpolar solutes in water. Contrary to expectations the effects of NaCl and Lil on the CMC of dioctanoyllecithin are not additive.  相似文献   

17.
The order to disorder transition of xanthan molecules in aqueous solutions has been studied using e.s.r. spectroscopy. Nitroxide spin-label was covalently attached to carboxyl groups on the xanthan side chains. The e.s.r. spectra obtained for aqueous spin-labelled xanthan solutions at varying ionic strengths contained both isotropic and anisotropic components at room temperature. The anisotropic component was attributed to the association of the side chains with the xanthan cellulosic backbone and was found to be present in greater proportions at increasing ionic strength. The spectra gradually changed with rising temperature and the proportion of anisotropic component decreased. This spectral change reflected the disruption of the side chain association with the backbone during the conformational change. Hysteresis effects were observed following sequential heating and cooling cycles suggesting that chain aggregation occurred.  相似文献   

18.
The effect of ion strength and pH value of the eluent on the determination of the lipophilicity of chlorhexidine was studied by reversed-phase thin-layer chromatography. The method has been improved by using various buffers: aqueous solutions of formic, acetic and propionic acids and their sodium salts in different ratios and in various concentrations. Stepwise regression analysis separated the effect of pH value, ion strength and acid type on the lipophilicity of chlorhexidine and proved that the ion strength exerted a higher impact than the pH value did. The effect of alkyl chain length of the acids was of secondary importance.  相似文献   

19.
Hyaluronate in cultured skin fibroblasts derived from patients with Werner's syndrome, who excrete large amounts of urinary hyaluronate, was investigated. The amount of hyaluronate secreted into the medium by Werner's fibroblasts was 2-3-times that of normal fibroblasts, whereas no difference in enzyme activities related to the degradation of hyaluronate was found. Werner's fibroblasts were then cultured in the presence of [3H]glucosamine, and the amount of [3H]hyaluronate and its chain lengths in the medium and matrix (trypsinate) fractions were compared with those of normal cells. No significant difference in the chain length of hyaluronate was observed between normal and Werner's fibroblasts. On the other hand, a significant increase of hyaluronate was found in the matrix fraction of Werner's fibroblasts when the cells reached confluency. In addition, a hyaluronate of small chain length was found in the matrix fraction of Werner's fibroblasts, although this was absent from that of normal cells. It was concluded that the constituents of the extracellular matrix of Werner's fibroblasts differed from those of normal cells, characterized by the presence of a large amount of hyaluronate and a relatively small hyaluronate chain.  相似文献   

20.
The viscosity behaviour of alginate-Cu2+-NaCl systems has been experimentally examined at various concentrations of cupric and sodium salts. Dependence of the intrinsic viscosity of alginate as a function of NaCl concentration is discussed to supplement the previous study which shows a similar behaviour to that found for other polyelectrolytes in aqueous solution in the presence of an added salt. The effects of sodium ions on the cupric association in cupric-induced alginate solutions were investigated by means of viscosity measurements. The mechanisms of complex formation in the presence of the simple added salt were studied. It was found that, at a given NaCl concentration, the viscosity of the mixture will pass through a maximum with increasing cupric concentration. The amounts of cupric cations corresponding to the maximum depends on the concentration of NaCl in the solution. Comparison of salt effects on the viscosity behaviour of alginate solutions during sol—gel transition reveals that an optimum NaCl concentration of 10−2 mol 1−1 exists where the viscosity of the mixture gives a maximum value at a certain cupric amount. This result indicates that salt effects play an important role in the sol—gel transition of the polyelectrolyte solutions. The observed phenomenon was interpreted in terms of conformational change of polyelectrolyte chain due to the addition of salt resulting in a different cross-linking mode in the system.  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