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.
Analysis of the mutual diffusion coefficient of hyaluronate reveals that it rapidly increases with increasing concentration or decreasing ionic strength. The mutual diffusion coefficients analyzed by boundary relaxation in the analytical ultracentrifuge by either Raleigh interference optics or absorption optics (through the use of fluorescein-labeled hyaluronate) yielded similar values. The theoretical treatment of the mutual diffusion coefficient has been analyzed in terms of experimentally measured intradiffusion coefficients and thermodynamic virial coefficients. Only approximate agreement between theory and experiment was found. The concept of formation of transient statistical network structures in semidilute solutions of hyaluronate was applied to evaluate a critical concentration at which network formation occurs. This has been discussed in relation to the marked decrease in the intradiffusion coefficient of hyaluronate with concentration. The formation of network structures in hyaluronate was found not to preclude the hyaluronate undergoing extremely rapid rates of mutual diffusion (with diffusion coefficients ~30 × 10?11 m2 s1) under conditions of relatively large initial chemical potential gradients. Measurements of the unidirectional flux of hyaluronate for nonzero gradients demonstrated their marked sensitivity to the magnitude of the concentration difference across the boundary. An experimental feature of the unidrectional diffusion coefficients of hyaluronate is that they may be analyzed purely in terms of mutual and intradiffusion processes. The backflux diffusion coefficient (describing the flux against the imposed concentration gradient) appeared identical with the intradiffusion coefficient. The analysis of the various sources of errors made in this study suggests that the magnitude of the diffusion coefficients measured may be regarded only as approximate.  相似文献   

2.
Laser correlation Spectroscopy was used to measure the mutual diffusion coefficient, D, of human cyanomethemoglobin (Fe+++:CN) at varying protein concentrations. These measurements were male at 20°C in a 0.1 M phosphate buffer solution at pH 7.0. For low protein concentrations we find D = (6.43 ± 0.26) × 10?7 cm2/S and that there is a near linear decrease from this value at higher concentrations. The linear relation between the diffusion coefficient and protein concentration allows us to deduce the value of the linear frictional volume fraction coefficient, Kf= 7.75. and to extrapolate to hemoglobin concentrations equivalent to that in the red blood cell where we estimate D = 4.25 × 10?7 cm2/s Various theoretical predictions of the dependence of the mutual diffusion coefficient on concentration are tested; we find that the generalized Stokes-Einstein relation can be made to fit our high concentration data if we assume a hard-sphere model and if we include a term involving a hydrodynamic interaction integral.  相似文献   

3.
A laboratory-made sample of the polysaccharide xylinan (acetan) has been further characterized with respect to (i) purity, (ii) molar mass and polydispersity, and (iii) gross conformation by a combination of hydrodynamic measurements (sedimentation velocity and equilibrium analytical ultracentrifugation, viscometry, and dynamic light scattering) in aqueous NaCl (I = 0.10 mol·L−1). Sedimentation velocity diagrams recorded using Schlieren optics revealed highly pure material sedimenting as a single boundary [so20.w = 9.5 ± 0.7) S; ks = (273 ± 112) mL/g]. The hypersharp nature of these boundaries is symptomatic of a polydisperse and highly nonideal (in the thermodynamic sense) system. Low speed sedimentation equilibrium in the analytical ultracentrifuge using Rayleigh interference optics and two different types of extrapolation procedure (involving point and whole-cell molar masses) gave a weight average molar mass Mw of (2.5 ± 0.5) × 10−6 g·mol−1 and also a second virial coefficient, B = (2.8 ± 0.7) × 10−4 mL·mol·g−2, both values in good agreement with those from light scattering-based procedures (Part II of this series). A dynamic Zimm plot from dynamic light scattering measurements gave a z-average translational diffusion coefficient Do20.w = (3.02 ± 0.05) × 10−8 cm2·s−1 and the concentration-dependence parameter kD = (370 ± 15) mL/g. Combination of so20.w with Do20.w via the Svedberg equation gave another estimate for Mw of ≅ 2.4 × 106 g/mol, again in good agreement. Both the Wales-van Holde ratio (ks/[η]) ≅ 0.4 (with [η] = (760 ± 77) mL/g) and the ρ-parameter (ratio of the radius of gyration from static light scattering to the hydrodynamic radius from dynamic light scattering) as ρ > 2.0 all indicate an extended conformation for the macromolecules in solution. These findings, plus Rinde-type simulations of the sedimentation equilibrium data are all consistent with the interpretation in terms of a unimodal wormlike coil model performed earlier. © 1996 John Wiley & Sons, Inc.  相似文献   

4.
The electrophoretic freeze-fracture electron microscopy method (Sowers, A.E. and Hackenbrock, C.R. (1984) Proc. Natl. Acad. Sci. USA 78, 6246–6250) for measuring the lateral diffusion coefficient of integral proteins was applied to a large population of spherical-shaped mitochondrial inner membranes. Membrane integral protein concentration was estimated by determining the intramembrane particle concentration. Analysis of the data reveals that: (a) the radii of the spherical inner membranes in the selected population ranged from 0.22 to 1.2 μm, (b) the intramembrane particle concentrations ranged from 2300 to 6400 per μm2, and (c) the calculated lateral diffusion coefficients of the intramembrane particles ranged from 1.3·10−10 to 3.35·10−9 cm2/s. The data clearly show a naturally occurring large range in protein concentration in the mitochondrial inner membrane and an inverse correlation of lateral diffusion coefficient with the membrane protein concentration. This study is the first to show that the lateral diffusion coefficient of integral proteins in a native membrane varies as the membrane protein concentration.  相似文献   

5.
The response of chickpea (Cicer arietinum L.) cv. KPG-59 to pre-sowing seed treatment with 28-homobrassinolide (HBR) and/or sodium chloride (NaCl) was investigated. The seeds imbibed in aqueous solution of 10−10 or 10−8 M of HBR for 8 h, resulted in an increase in the values for most of the characteristics of shoot and root at 90-day stage and seed yield, at harvest. The plants resulting from the seeds soaked in HBR (10−8 M) possessed 23% and 31% higher leaf nitrate reductase (E.C. 1.6.6.1) and carbonic anhydrase (E.C. 4.2.2.1) activities, 34% more dry mass, 30% higher nodule number, 31% and 18% more nodule fresh and dry mass, compared with water soaked, control. Leghaemoglobin content and nitrogenase activity (E.C. 1.7.99.2) were 28% and 30% higher while nodule nitrogen and carbohydrate contents decreased by 5% and 6%, compared with the control. Moreover, seed yield increased by 26% over the control, at harvest. The values for all the above characteristics declined significantly, in the plants raised from the seeds soaked in NaCl. However, this ill effect was overcome, if NaCl treatment was given before or after HBR treatment.  相似文献   

6.
The hemoglobins of three snake species: Liophis miliaris, Bothrops alternatus and Boa constrictor present a single ATP binding site per tetramer. The ATP association constant values for the deoxyhemoglobins at pH 7.5 were about KD ≅ 106 M−1 (107 M−1 for B. contrictor), three to four orders of magnitude higher than the respective values for oxyhemoglobin of about KO ≅ 102 M−1. The deoxyhemoglobin constant values markedly decrease as a function of pH, becoming, at pH 8.5, about KD ≅ 103 M−1 whereas for the oxyhemoglobin the constants remain of about the same, KO ≅ 102 M−1, at the pH range studied. The high ATP binding affinity constants, compared to those of human hemoglobin A, were explained from a molecular structural standpoint, considering L. miliaris hemoglobin, whose complete primary sequence is known. Two distinct amino acid residue differences were found in the β-chain, one being Trp (NA3) (more hydrophobic) in the snake hemoglobin which substitutes the Leu (NA3) in human hemoglobin, and the second being Val 101 β (G3) instead of Glu 101 β (G3). The substitutions could provide an un-neutralized, positively charged, residue Lys-104β and, taking into account its high pK value, the pH dependence of ATP binding affinity for the snake hemoglobin would originate from pH-dependent ionization of phosphate groups of the allosteric effector. The physiological implications of the high ATP binding constant, as well as the possible protective role of the nucleotide binding against the effect of high environmental temperatures on the oxygen dissociation curves, are discussed.  相似文献   

7.
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   

8.
Abstract

A high molecular weight ‘cryogel’ was obtained as insoluble complexes by cold incubation at near-freezing temperatures from heparinized plasma of patients with rheumatoid arthritis. After the cryogel was solubilized at 37°C, 1:1 complex of fibrinogen and fibronectin was purified at room temperature by affinity chromatography on a gelatin-Sepharose 4B. Hydrodynamic properties of the complex were investigated as a function of temperature and NaCl concentration using a dynamic light scattering. The diffusion coefficients of the complex at 20°C decreased with increasing of NaCl concentration as free fibronectin. The complex appears to be a more compact form at low ionic concentration, which is associated with conformational changes of fibronectin. The diffusion coefficient of the complex at 20°C in 0.05 M Tris- HCl(pH7.4) containing 0.5 M NaCl was estimated as 8.5× 10?8 cm2s?1. The complex did not dissociate over the temperature range from 20 to 37°C. The diffusion coefficients of the complex decreased significantly at 12°C and 40°C. The thermal denaturation of fibrinogen molecule in the complex was observed at 40°C. The CONTIN analysis of the light scattering data showed that the complex associated to form higher aggregates at 15°C, but not at near- freezing temperature. The equilibrium between the complex and higher aggregates appeared reversible.  相似文献   

9.
The kinetics of uptake and loss of Na+ have been studied using radioisotopic tracer techniques in cells of the cyanobacterium Synechocystis PCC6714 exposed to hyperosmotic stress. Cells transferred from a fresh-water-based medium to NaCl at 300–1000 mmol·dm−3 showed net Na+ uptake during the first 2 min following transfer, with the intracellular Na+ level at 2 min increasing as a direct function of the external NaCl concentration. Further incubation of cells in low-level hypersaline media (350–500 mmol · dm−3 NaCl) led to a marked reduction in cell Na+ within 20 min, indicating an efficient active Na+ extrusion system. In contrast, cells maintained in more extreme hypersaline media showed little (750 mmol · dm−3 NaCl) or no (1000 mmol · dm−3 NaCl) net Na+ extrusion following upshock. Cells grown in a saline medium (with NaCl at 500 mmol · dm−3 showed a greatly reduced net Na+ uptake after 2 min in media containing higher levels of NaCl. However, net Na+ uptake was also observed when these cells were downshocked to media containing 50–200 mmol · dm−3 NaCl. This is the first demonstration of downshock-induced Na+ accumulation in a microbial cell. Time-courses for Na+ extrusion in cells downshocked from 500 mmol · dm−3 to 100 mmol · dm−3 NaCl were similar to those for cells upshocked from freshwater to 500 mmol · dm−3 NaCl, requiring approximately 30 min to reach their lowest values. Net Na+ extrusion in upshocked cells was found to be markedly sensitive to the external K+ concentration, with limited net Na+ extrusion in the absence of external K+ and maximal reductions in cell Na+ in media containing K+ at 1–10 mmol · dm−3. Temperature was also shown to affect uptake and loss of cell Na+ during upshock: cells maintained at 5°C showed no capacity for net Na+ extrusion, while higher temperatures (up to 35°C) led to a progressive reduction in the amount of cell Na+ at 2 min following upshock and also in the rate of net Na+ extrusion after this time.  相似文献   

10.
Static and dynamic light-scattering measurements are reported on zinc-insulin at room temperature (21 ± l°C) and pH = 6.88 in 0.1M NaCl aqueous solution. The experiments were performed at very low concentration, in the range 0.12 × 10?4 to 0.90 × 10?4 g cm?3. Within experimental error, we find no evidence for a critical micellar concentration in this system. The aggregation phenomenon starts immediately after preparation of the solutions, and takes several days to come to stable equilibrium. The concentration dependence of the diffusion coefficients, D z, = Do (1 — kDC), is negative, and kD was observed to decrease as a function of time, while the aggregate size was found to increase. The equivalent concentration coefficient, ?2BM W, obtained from static light scattering, showed a similar behavior, and, within experimental error, was found to be numerically equal to kD. From the relation found between the diffusion coefficient at infinite dilution and the molecular weight of the aggregates, log D0 = ?0.240 log M w ? 5.077, we deduce that the insulin aggregates are compact structures with a characteristic radius of 0.71 Å/(dalton)1/3, surrounded by a hydration layer of a thickness of 8.0 Å. The equilibrium aggregation number is approximately 10.  相似文献   

11.
The mixed cultures has been isolated from industrial saline wastewater contaminated with chromium(VI), using enrichment in the presence of 50 mg l−1 chromium(VI) and 4% (w/v) NaCl at pH 8. In this study, the molasses (M) medium was selected a suitable medium for the effective chromium bioaccumulation by the mixed cultures. Eleven pure isolates obtained from mixed cultures and some of them showed high bioaccumulation in the M media containing about 100 mg l−1 chromium(VI) and 4% NaCl. The strain 8 (99.3%) and 10 (99.1%) were able to bioaccumulate more efficient than the mixed culture (98.9%) in this media. But the highest specific Cr uptake was obtained by the mixed cultures followed by strain 8 and 10 with 56.71, 33.14 and 21.7 mg g−1, respectively. Bioaccumulation of chromium(VI) ions by the strain 8 growing in the media with chromium(VI) and NaCl was studied in a batch system as a function of initial chromium(VI) (86.6–547.6 mg l−1) and NaCl (0, 2, 4, 6% w/v) concentrations. During all the experiments, the uptake yield of the strain 8 was highly affected from NaCl concentrations in the medium at high initial chromium(VI) concentrations. But at low chromium(VI) concentration, strain 8 was not affected from NaCl concentrations in the medium. The maximum uptake yield were obtained in the M media with 2% NaCl as 98.8% for 110.0 mg l−1, 98.6% for 217.1 mg l−1, 98.6% for 381.7 mg l−1 and 98.2% for 547.6 mg l−1 initial chromium(VI) concentrations. The strain 8 tolerated a 6% (w/v) NaCl concentration was able to bioaccumulate more than 95% of the applied chromium(VI) at the 97.6–224.4 mg l−1 initial chromium(VI) concentrations. The results presented in this paper was shown that these pure and mixed cultures might be of use for the bioaccumulation of chromium(VI) from saline wastewater.  相似文献   

12.
The collision rates between spin-labelled valeric acid in water, and between the corresponding mixed-chain, spin-labelled phosphatidylcholine in water-methanol mixtures, and also between spin-labelled phosphatidylcholine monomers and micelles in water have been determined from the spin-spin broadening of the electron spin resonance spectrum. In each case the second order rate constants are consistent with a diffusion-controlled process. For spin-labelled valeric acid in water the translational diffusion coefficient at 20°C is 3.4 · 10−6 cm2 · s−1, and for spin-labelled phosphatidylcholine varies between 2.3 · 10−6 and 3.8 · 10−6 cm2 · s−1 within the range 44 to 88 wt% methanol. The spin-labelled phosphatidylcholine monomer diffusion coefficient in water at 20°C is 2.4 · 10−6 cm2 · s−1, deduced from the monomer-micelle association rate, with an activation energy of 4.0 kcal · mol−1. The much slower on-rates for association of lipid monomers with phospholipid bilayer vesicles reported in the literature, therefore indicate that incorporation into bilayers is not a diffusion-controlled process.  相似文献   

13.
Pakistan’s most of the land is less productive or no productivity at all due to erosion and salinity of the soil, which can be utilized to develop fisheries. The project, “Survival, growth and body composition of Cyprinus carpio under different salinity regimes” was undertaken in two phases. In the first phase susceptibility of Cyprinus carpio at four salinity levels in triplicate within 0–10 g L−1NaCl for 96 h in each aquarium was checked after one week acclamation at 0 g L−1, 2 g L−1 and 4 g L−1 NaCl. LC50 values varied from 7.67 to 10.65 g L−1 after 96 h for C. carpio. Percentage mortality of the fish and important water quality parameters after every 12 h were observed for a period of 96-h. Probit analysis showed that 96-h LC50 values ranged from 7.67 to 10.65 g L−1. During experimental period aquaria water temperature ranged from 29.6 to 33.7 °C, pH values fluctuated between 7.8 and 9.7, Electrical conductivity values ranged from 2.40 to 20.13 dSm−1 and Dissolved oxygen ranged between 2.23 and 10 mg L−1. Sub-lethal salt concentration i.e. 0 g L−1 to 3 g L−1 NaCl upto 40 days showed that growth of C. carpio decreased with the increase of water salinity levels and ceased at 4 g L−1 salinity and increase in salinity have negatively affected hematological parameters.  相似文献   

14.

In the domain of production of flavor delivery systems based on carbohydrate matrices, Hot Melt Extrusion (HME) is one of the processes used which implies a relatively slow water evaporation step compared to Spray Drying or other common processes. The HME process of generating liquid syrup is stable only at industrial scales. The drying kinetic associated to the syrup formation is followed by collecting the vapored water. This experiment is exploited to estimate the diffusion coefficient of liquid water in a continuously concentrating carbohydrate melt ranging from 30.4 to 10.2%w/w water content and at temperatures ranging from 100 to 116 °C. A combination of Fick, Arnold and Vogel-Tammann-Fulcher models was used to derive the viscosity of the solution, the gradient thickness at the surface of the stirred solution and the diffusion coefficient of water in the liquid under boiling conditions. It is here demonstrated that under continuous boiling conditions, the water concentration gradient depth from the surface of the solution has a dimension that is comparable to the size of the hydrated carbohydrate molecules and the diffusion coefficient of water in boiling carbohydrate solutions ranges from 5.10−9 cm.s−2 for low temperature and high water content to 10−11 cm.s−2 for high temperature and low water content.

  相似文献   

15.
Delivering charged antibiotics to the intervertebral disc is challenging because of the avascular, negatively charged extracellular matrix (ECM) of the tissue. The purpose of this study was to measure the apparent diffusion coefficient of two clinically relevant, charged antibiotics, vancomycin (positively charged) and oxacillin (negatively charged) in IVD. A one-dimensional steady state diffusion experiment was employed to measure the apparent diffusion coefficient of the two antibiotics in bovine coccygeal annulus fibrosus (AF) tissue. The averaged apparent diffusion coefficient for vancomycin under 20% compressive strain was 7.94 ± 2.00 × 10−12 m2/s (n = 10), while that of oxacillin was 2.26 ± 0.68 × 10−10 m2/s (n = 10). A student’s t-test showed that the diffusivity of vancomycin was significantly lower than that of oxacillin. This finding may be attributed to two factors: solute size and possible binding effects. Vancomycin is approximately 3 times larger in molecular weight than oxacillin, meaning that steric hindrance likely plays a role in the slower transport. Reversible binding between positive vancomycin and the negative ECM could also slow down the rate of diffusion. Therefore, more investigation is necessary to determine the specific relationship between net charge on antibiotic and diffusion coefficients in IVD. This study provides essential quantitative information regarding the transport rates of antibiotics in the IVD, which is critical in using computational modeling to design effective strategies to treat disc infection.  相似文献   

16.
This work evaluates the influences of five parameters (pH, PEG molecular mass, PEG concentration, concentration of buffer K2HPO4–KH2PO4 and NaCl concentration) on xylanolitic complex partitioning produced by P. janthinellum in aqueous two-phase systems, using a 25 factorial experimental design. A mathematical model to quantify the influence of these parameters was attained and statistically tested. The optimum point for total protein extraction was obtained under the following conditions: pH 7.0, PEG 10 000, 3.67% PEG, 10% potassium phosphate and 12.4% NaCl. The partition coefficient (K) value experimentally obtained was 5.25 and that predicted by the model was 5.89.  相似文献   

17.
Dynamic and static light scattering, CD, and optical melting experiments have been conducted on M13mp19 viral circular single-strand DNA as a function of NaCl concentration. Over the 10,000-fold range in concentration from 100 microM to 1.0 M NaCl, the melting curves and CD spectra indicate an increase in base stacking and stability of stacked regions with increased salt concentration. Analysis of dynamic light scattering measurements of the single-strand DNA solutions as a function of K2 from 1.56 to 20 X 10(10) cm-2 indicates the collected autocorrelation functions are biexponential, thus revealing the presence of two decaying dynamic components. These components are taken to correspond to (1) global translational motions of the molecular center of mass and (2) motions of the internal molecular subunits. From the evaluated relaxation rates of these components, diffusion coefficients D0 and Dplat are determined. The center of mass translational diffusion coefficient D0, varies in a nonmonotonic manner, by 10%, from 3.75 X 10(-8) to 3.39 X 10(-8) cm2/s over the NaCl concentration range from 100 microM to 1.0 M. Likewise, the radius of gyration RG, obtained from static light scattering experiments, varies by 15% from 699 to 830 A over the same NaCl range Dplat, the diffusion coefficient of the internal subunits, displays a different dependence on the NaCl concentration and decreases, by nearly 22% in a titratable fashion, from 12.46 X 10(-8) to 10.26 X 10(-8) cm2/s, when the salt is increased from 100 microM to 1.0 M. A semiquantitative interpretation of these results is provided by analysis of the light scattering data in terms of the circular Rouse-Zimm chain. Rouse-Zimm model parameters are estimated from the experimental results, assuming the circular chains are composed of a fixed number of Gaussian segments, N + 1 = 15. The rms displacement of the internal segments, b, is estimated to be the smallest (442 A) in 100 mM NaCl. Increases of b to 467 A in 100 microM and 524 A in 1.0 M NaCl are observed. Meanwhile, the hypothetical friction factor of the internal subunits, f, progressively increases as the NaCl concentration is raised. It is inferred from the evaluated Rouse-Zimm model parameters that both the static flexibility of the circular chain and diffusive displacements of the internal subunits decrease with increases in NaCl concentration from 100 mM to 1.0 M.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
《Process Biochemistry》2014,49(1):33-37
The ectoine-excreting bacterial strain of Halomonas salina was employed in the co-production of poly-β-hydroxybutyrate (PHB) and ectoine (Ect) during a fermentation process (PHB/Ect co-production). An efficient PHB/Ect co-production process was carried out at low NaCl concentration (30 g L−1). It was established using 1H Nuclear Magnetic Resonance spectroscopy that H. salina produces PHB. The effects of the NaCl concentration, the initial C/N ratio, the phosphate concentration and mixed carbon sources were investigated with respect to PHB/Ect co-production. The PHB/Ect co-production system comprised growing and non-growing cell phases and was developed with NaCl concentration of 30 g L−1. The optimal conditions for PHB/Ect co-production by the ectoine-excreting strain of H. salina were 30 g L−1 NaCl, with an initial C/N ratio of 15, an initial phosphate concentration of 12 g L−1 and mixed carbon sources of 55 g L−1 glucose and 25 g L−1 monosodium glutamate. Using a PHB/Ect co-production system with growing and non-growing cell phases prevents the inhibition of PHB synthesis by high concentration of NaCl and significantly reduces ectoine degradation. PHB and ectoine concentrations as high as 35.3 g L−1 and 8.6 g L−1, respectively, were achieved. The efficient co-production of PHB and ectoine at a low NaCl concentration has been realised.  相似文献   

19.
Sorbitan trioleate was modified with Cibacron Blue F-3GA (CB) to create an affinity surfactant and to form affinity-based reverse micelles in n-hexane. The partitioning equilibria and the extraction kinetics of lysozyme and bovine serum albumin (BSA) were then examined. The solubilization capacity of the reverse micellar system for lysozyme increased linearly with increasing the CB concentration from 0.1 to 0.5 mmol L−1. In contrast, the capacity for BSA at 0.5 mmol L−1 of coupled CB was only about one-fifth that for lysozyme. It indicates a strong steric hindrance effect of the micelles for the high molecular mass protein. The overall volumetric mass transfer coefficient of lysozyme in the forward extraction increased from 0.43 × 10−3 to 1.25 × 10−3 s−1 with increasing CB concentration from 0.1 to 0.5 mmol L−1. Due to the high molecular mass of BSA, its volumetric mass transfer coefficient in the forward extraction was only one-sixth that of lysozyme. The ratio of the coefficient in the back extraction to that in the forward extraction was less than 0.03, much lower than those in other micellar systems. It indicates that the interfacial resistance in this system was severer than in others.  相似文献   

20.
Self-association of hyaluronate segments in aqueous NaCl solution   总被引:3,自引:0,他引:3  
The potential for self-association by hyaluronate (HA) chains in 0.15 M NaCl was investigated, using low molecular weight HA segments as a model system. HA segments were derived from the polymer by controlled enzymatic digestion, and purified by gel filtration chromatography. Seven samples of narrow molecular weight distribution were analyzed by sensitivity-enhanced polyacrylamide gel electrophoresis, and found to have the following weight-average numbers of repeating disaccharide units: A, 90; B, 51; C, 38; D, 31; E, 23; F, 18; G, 13. The segment preparations were studied in 0.15 M NaCl by capillary viscometry, low angle laser light scattering, and circular dichroism spectroscopy. The data indicate concentration-dependent intermolecular association of short segments, and a capability for intramolecular association (hairpin formation) by larger HA segments.  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