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 13 毫秒
1.
To gain new insights into the interaction of proteins and disaccharides, we investigated the hydrodynamic radii, , of lysozyme molecules in solution and in a ternary protein‐sugar‐water system by PFG‐NMR. Our approach is based on the assumption that the anhydrobiotic properties of disaccharides like trehalose are based on aggregation of sugar molecules to the proteins, i.e., accumulation of sugar molecules close to the protein, and that this process can be investigated by the experimentally detectable value of the protein. The Rh values are calculated from the experimentally determined diffusion coefficients and the application of a viscosity correction using the inert molecule dioxane as an internal viscosity reference. The experiments were performed as a function of sugar concentration, the overall particle concentration and the pH value. We investigated the disaccharides trehalose and sucrose, mainly for the reason that trehalose has well know cryptobiotic properties while sucrose, which is similar in size and structure, lacks these properties. The results show the formation of a protective sugar shell around the proteins over a wider range of concentrations and pH values in the case of trehalose.  相似文献   

2.
Depolymerization of N-succinyl-chitosan by hydrochloric acid   总被引:2,自引:0,他引:2  
N-Succinyl-chitosan (1) was depolymerized with 7.5 M aqueous HCl at room temperature or 3.3 M aqueous HCl at 40 degrees C and the molecular weights (MW) of the products were determined by size-exclusion chromatography-multi angle light scattering (SEC-MALS) and their viscometric features were investigated. The intrinsic viscosity ([eta]) obtained at the concentration of 0.1-0.3% (w/v) in saline showed a linear relationship between log[eta] and log MW, which provided the coefficients in the Mark-Houwink equation.  相似文献   

3.
The dynamic mechanical behaviour of a series of concentrations of kappa-carrageenan (KC; 0.35-1.6% w/w) and iota-carrageenan (IC; 0. 2-1% w/w) in 0.2 M NaI has been investigated. The flow behaviour of KC within the concentration range 0.004-0.8% (w/v) was also described. The high intrinsic viscosity of KC in 0.2 M NaI (23.4 dl g(-1)) and the great increase in viscosity with increasing concentration, in comparison with linear flexible polysaccharides, is well consistent with the stiffness of KC helices in NaI. The variation of the 'zero-shear' specific viscosity of KC in 0.2 M NaI with the degree of space-occupancy (c[eta]) displays two critical concentrations at c* approximately 0.09% w/v (c*[eta] approximately 2) and c** approximately 0.4% w/v (c**[eta] approximately 10). Different viscoelastic behaviours were exhibited from the liquid-like to the solid-like depending upon the type of carrageenan. From the application of the time-temperature superposition, classical frequency-temperature master curves could be obtained for KC, but not for IC. Moreover, for KC, a concentration-frequency master curve could be constructed for the concentrations below 1.5%, indicating a 'solution-like' behaviour in this entire concentration range, although systems above 0.8% were visually gel-like. It is proposed that the rigidity of the KC helices is responsible for the slow relaxation rates of the gel-like samples. At higher concentrations (beyond 1.6%) a frequency-temperature superposition was no longer possible. In contrast to KC, IC behaved as a typical viscoelastic gel with a very weak frequency dependence of the storage modulus at all temperatures. This indicates the existence of associations beyond simple entanglements for IC. Creep experiments performed at higher carrageenan concentrations in 0.1 M NaI further corroborated the differences in the viscoelastic behaviour between KC and IC.  相似文献   

4.
The flexibility and curvature of duplex DNAs containing mismatched sites have been monitored as a function of temperature. The diffusion coefficients are dependent on the flexibility and the curvature of the DNA, and these have been determined by NMR-based methods. The diffusion coefficients, D, depend on a Boltzmann term and the viscosity of the solvent, eta, which is also temperature dependent. To analyze the temperature dependence of the diffusion results, the shape function, S(f) = etaD/T, is used. The shape functions do not have the viscosity and temperature dependence of the diffusion coefficients. The presence of mismatched sites significantly enhances the shape function of duplex DNA at all temperatures examined. The observed increases in the shape functions are attributed to the mismatched sites acting as localized flexible joints. The results on the temperature dependence of the shape functions, the optical absorbance, and the proton chemical shifts indicate that local melting at, and adjacent to, mismatched site occurs at a lower temperature than the overall melting of the duplexes. The localized melting gives rise to a considerable increase in the shape function. The contribution of the curvature of the mismatched sites to the enhanced diffusion has been examined. A DNA with mismatches that are in phase with respect to the helical repeat and a DNA which has the mismatches out of phase with respect to the helical repeat have been examined. The results indicate that mismatched sites have modest curvature.  相似文献   

5.
6.
Data are presented on the rheological and hemolytic behavior of whole human blood as it ages while stored at 4 degrees C (as in blood banking practice) up to 26 days. The viscometric properties of steady shear viscosity eta and oscillatory (complex) viscosity eta * = eta' - i eta" reported over ranges of shear rate gamma and radian frequency omega of 33 less than gamma less than 4130 s-1 and 1.5 less than omega less than 48 s -1; data on autologous plasma are given for reference. The Cox-Merz relation, eta (gamma) = [eta *(omega)] omega = gamma, is found to be a good approximation, with eta greater than or equal to [eta *], over the range studied. Release of hemoglobin (Hgb) and lactate dehydrogenase (LDH) into the plasma during shearing is tracked as a function of time for 30 min, and its sensitivity to gamma magnitude is measured. Bloods from four different donors are studied, with primary attention given to one (SSR). For all bloods, the release of both Hgb and LDH increases with storage age, but differences in such aging characteristics between different bloods can be substantial (even when rheological properties are identical). A post-shear incubation at 4 degrees C for one day shows no enhancement of plasma Hgb and LDH levels beyond those expected from normal aging after the shearing experience, demonstrating the absence of significant delayed-action effects as a consequence of shearing trauma.  相似文献   

7.
Defining the self-association state of a molecule in solution can be an important step in NMR-based structure determination. This is particularly true of peptides, where there can be a relatively small number of long-range interactions and misinterpretation of an intermolecular NOE as an intramolecular contact can have a dramatic influence on the final calculated structure. In this paper, we have investigated the use of translational self-diffusion coefficient measurements to detect self-association in aqueous trifluoroethanol of three peptides which are analogues of the C-terminal region of human neuropeptide Y. Experimentally measured diffusion coefficients were extrapolated to D0, the limiting value as the peptide concentration approaches zero, and then converted to D20,w, the diffusion coefficient after correction for temperature and the viscosity of the solvent. A decrease in D20,w of about 16% was found for all three peptides in aqueous TFE (30% by volume) compared with water, which is in reasonable agreement with the expected decrease upon dimerisation, the presence of which was indicated by sedimentation equilibrium measurements. Apparent molecular masses of these peptides in both solutions were also calculated from their diffusion coefficients and similar results were obtained. Several potential internal standards, including acetone, acetonitrile, dimethylsulfoxide and dioxane, were assessed as monitors of solution viscosity over a range of trifluoroethanol concentrations. Compared with independent measurements of viscosity, acetonitrile was the most accurate standard among these four. The practical limitations of a quantitative assessment of peptide self-association from translational diffusion coefficients measured by PFGNMR, including the calculation of apparent molecular mass, are also discussed.  相似文献   

8.
The diffusivity of several monosaccharides and disaccharides in calcium alginate gels was determined using a specially designed diaphragm cell. The diffusion coefficients of the tested sugars are 4 to 18% smaller in alginate gel than in water and, with the exception of fructose, this difference increases with increasing sugar molecular weight. Also the position of the carbonyl group seems to be determined in the value of the diffusion coefficient - ketoses have lower diffusion coefficients than aldoses.  相似文献   

9.
The rotational mobility of the phosphate translocator from the chloroplast envelope and of lipid molecules in the membrane of unilamellar azolectin liposomes has been investigated. The rotational dynamics of the liposome membrane were investigated by measuring the rotational diffusion of eosin-5-isothiocyanate(EITC)-labeled L-alpha-dipalmitoylglycerophosphoethanolamine (Pam2 GroPEtn) in the lipid phase of the vesicles, either in the presence or absence of the reconstituted phosphate translocator. The temperature dependence of the anisotropy decay showed that above 25 degrees C the main contribution to the anisotropy decay was caused by uniaxial anisotropic rotation of the labelled lipid molecules around the axis normal to the membrane plane. The rate of rotation of the labelled lipid molecules was strongly dependent on the viscosity of the medium (eta 1). Extrapolation to eta 1 = 0 Pa.s yielded a correlation time of phi = 20 +/- 5 ns, t = 30 degrees C, for lipid rotation with respect to the membrane normal. The rotational diffusion coefficient of the lipid molecules was calculated to be Dr = 2.0 x 10(9) rad2.s-1 and the apparent microviscosity in the vesicle membrane, as derived from the rotational correlation time, was eta 2 approximately 12 mPa.s. The rotational correlation time of the phosphate translocator in the membrane was only slightly dependent on the viscosity of the medium. The temperature dependence of the protein rotation also indicated that the rotation of the protein in the membrane was largely restricted and occurred mainly about the axis normal to the membrane plane. Measurements at a medium viscosity of eta 1 = 1 mPa.s yielded a value of phi r approximately 450 ns corresponding to Dr = 8.8 x 10(7) rad2.s-1 for protein rotation with respect to the membrane normal. From this value and the data of the lipid rotation, the cross-sectional area of the protein part embedded in the membrane was calculated to be approximately 9 nm2. This cross-sectional area is large enough to include at most 14 membrane-spanning helices. Our results also indicated that at lipid/protein molar ratios greater than or equal to 1.5 x 10(4): 1 aggregation occurred in the model membranes below 30 degrees C. However, above 30 degrees C and at a high dilution of the protein in the membrane it appeared that the membrane viscosity monitored by lipid and protein rotational diffusion were identical.  相似文献   

10.
We studied the effects of calcium ion concentration on the temperature dependence of rheological behavior of human red blood cells (RBCs) and concentrated hemoglobin solutions. Our previous study (G. M. Artmann, C. Kelemen, D. Porst, G. Büldt, and S. Chien, 1998, Biophys. J., 75:3179-3183) showed a critical temperature (Tc) of 36.4 +/- 0.3 degrees C at which the RBCs underwent a transition from non-passage to passage through 1.3 microm micropipettes in response to an aspiration pressure of -2.3 kPa. An increase in intracellular Ca2+ concentration by using the ionophore A23187 reduced the passability of intact RBCs through small micropipettes above T(c); the micropipette diameter needed for >90% passage increased to 1.7 microm. Viscometry of concentrated hemoglobin solutions (45 and 50 g/dl) showed a sudden viscosity transition at 36 +/- 1 degrees C (Tc(eta)) at all calcium concentrations investigated. Below Tc(eta), the viscosity value of the concentrated hemoglobin solution at 1.8 mM Ca(2+) was higher than that at other concentrations (0.2 microM, 9 mM, and 18 mM). Above Tc(eta), the viscosity was almost Ca2+ independent. At 1.8 mM Ca2+ and 36 +/- 1 degrees C, the activation energy calculated from the viscometry data showed a strong dependence on the hemoglobin concentration. We propose that the transition of rheological behavior is attributable to a high-to-low viscosity transition mediated by a partial release of the hemoglobin-bound water.  相似文献   

11.
Nectar is an essential resource for bumblebees and many other flower-visiting insects. The main constituents of nectar are sugars, which vary in both composition and concentration between plant species. We assessed the influence of sugar concentration, sugar solution viscosity and sugar solution composition on the imbibition and energy intake rate of bumblebees, Bombus impatiens Cresson (Hymenoptera: Apidae). To do this, we measured their rate of solution intake for 49 different sugar solution treatments, which varied in both sugar composition and concentration. In general, the imbibition rates of bumblebees were found to increase with increasing sugar concentration, probably due to their preference for high sugar concentrations, up to a concentration of 27% (w/w), at which point solutions reached a threshold viscosity of approximately 1.5–1.6 mPa.s. Above this threshold, the increasing viscosity of the solutions physically inhibited the imbibition rates of bees, and imbibition rate began to decrease as the concentration increased. Nevertheless, bumblebee energy intake rate increased with increasing concentration up to about 42–56%. Although we found that sugar solution composition had an impact on both imbibition and energy intake rate, its effect was not as straightforward as that of sugar concentration and viscosity.  相似文献   

12.
The Smoluchowski equation for the bimolecular collision rate of dissolved oxygen molecules with spin labels yielded values for the diffusion constant of oxygen in water that are in agreement with the Stokes-Einstein equation (D infinity T/eta, where eta is the macroscopic viscosity) and with published values obtained by conventional methods. Heisenberg exchange at an interaction distance of 4.5 A occurs with a probability close to one for each encounter. In mixed hydrocarbons (olive oil, paraffin oils) and sec-butyl benzene, D infinity (T/eta)rho, where rho lies between 0.5 and 1. Oxygen diffuses in the hydrocarbons between 10 and 100 times more rapidly than predicted from the macroscopic viscosity. Similar results would be expected for diffusion of oxygen in model and biological membranes. Parallel measurements of rotational diffusion of the spin labels show little correlation with measurements of translational diffusion of oxygen. Dipolar interactions between spin labels and oxygen appear negligible except in the limit of highest viscosities.  相似文献   

13.
Isothermal-isobaric molecular dynamics simulations are used to calculate the specific volume of models of trehalose and three amorphous trehalose-water mixtures (2.9%, 4.5% and 5.3% (w/w) water, respectively) as a function of temperature. Plots of specific volume versus temperature exhibit a characteristic change in slope when the amorphous systems change from the glassy to the rubbery state and the intersection of the two regression lines provides an estimate of the glass transition temperature T(g). A comparison of the calculated and experimental T(g) values, as obtained from differential scanning calorimetry, shows that despite the predicted values being systematically higher (about 21-26K), the trend and the incremental differences between the T(g) values have been computed correctly: T(g)(5.3%(w/w))相似文献   

14.
We investigated the effects of temperature, water activity (a(w)), and syrup film composition on the CFU growth of Wallemia sebi in crystalline sugar. At a high a(w) (0.82) at both high (20 degrees C) and low (10 degrees C) temperatures, the CFU growth of W. sebi in both white and extrawhite sugar could be described using a modified Gompertz model. At a low a(w) (0.76), however, the modified Gompertz model could not be fitted to the CFU data obtained with the two sugars due to long CFU growth lags and low maximum specific CFU growth rates of W. sebi at 20 degrees C and due to the fact that growth did not occur at 10 degrees C. At an a(w) of 0.82, regardless of the temperature, the carrying capacity (i.e., the cell concentration at t = infinity) of extrawhite sugar was lower than that of white sugar. Together with the fact that the syrup film of extrawhite sugar contained less amino-nitrogen relative to other macronutrients than the syrup film of white sugar, these results suggest that CFU growth of W. sebi in extrawhite sugar may be nitrogen limited. We developed a secondary growth model which is able to predict colony growth lags of W. sebi on syrup agar as a function of temperature and a(w). The ability of this model to predict CFU growth lags of W. sebi in crystalline sugar was assessed.  相似文献   

15.
The flocculating activity of an exopolysaccharide, p-KG03, produced by a marine dinoflagellate Gyrodinium impudicum KG03 was investigated. The p-KG03 was a highly sulfated exopolysaccharide that showed strong antiviral activity against encephalomyocarditis virus (EMCV) and immunostimulating activity by NK cell activation. For the industrial applications of p-KG03, as the bioflocculant agent, p-KG03 showed that more than 90% of the flocculating activity in kaolin suspension occurred at concentrations of 0.5 mg/l with the maximum at 1.0 mg/l. However, flocculation decreased from 2.5 mg/l. The flocculation rate increased linearly with concentration and was higher than that observed in commercial products such as polyacrylamide (approximately 1.0 mg/l) or zooglan (approximately 3.0 mg/l). The p-KG03 was an effective flocculant under acidic conditions (pH 3-6) and over a wide temperature range (4-90 degrees C). The presence of cations did not enhance flocculating activity. The average molecular mass, as determined by gel filtration chromatography, was about 1.87 x 10(3) KDa. Galactose was the main sugar in p-KG03, which also contained uronic acid (2.9%, w/w) and sulfate groups (10.3%, w/w). The infrared spectrum of p-KG03 showed absorption bands of carboxylate groups. Thermogravimetric analysis indicated a degradation temperature (T(d)) of 250 degrees C. Several other properties of p-KG03 such as intrinsic viscosity, the rheological behavior, consistency index (k) and flow behavior index (eta) were also studied.  相似文献   

16.
Using 5 samples of well-purified Na-gellans (Na-gellans G1-G5, weight-average molar mass M(w) = 120 x 10(3)-32 x 10(3) at 40 degrees C), the effects of molar mass on the coil-to-double-helix transition in aqueous solutions with 25 mM NaCl were studied by light scattering and circular dichroism (CD) measurements, viscometry, and differential scanning calorimetry (DSC). From the temperature dependence of M(w), molar ellipticity at 201 nm [theta]201, intrinsic viscosity [eta], and DSC exothermic curves, it was found that the coil-to-double-helix transitions for G1-G5 samples took place at almost the same temperature. The [eta] and M(w) obtained in the temperature range from 40 to 25 degrees C can be explained by a simple coil/double-helix equilibrium model using the double-helix contents determined from CD data. The van't Hoff's transition enthalpy deltaH(vH) of Na-gellans depended on M(w). It is concluded that the coil-to-double-helix transitions of Na-gellans are all-or-none type transitions, and are accelerated with increasing M(w).  相似文献   

17.
The hydrodynamic parameters of the major protein fraction, viz. arachin from groundnut, alpha-globulin from sesame seed, brassin (M) from mustard seed and helianthinin from sunflower seed, have been determined in a single solvent system (0.05 M Tris-HCl buffer, pH 7.5 containing 0.5 M sodium chloride): sedimentation coefficient (s0(20,w)) and diffusion coefficient (D0(20,w)) by analytical ultracentrifugation, intrinsic viscosity [eta] by Ostwald viscometry and partial specific volume (V) by densimetry. The molecular weights (M) of the four proteins, calculated using the sedimentation-viscosity and sedimentation-diffusion coefficient methods, were found close to each other. The values have been compared with those in the literature and the reasons for discrepancies have been discussed.  相似文献   

18.
Y Takano  A Sakanishi 《Biorheology》1988,25(1-2):123-128
To consider the effects of the viscoelasticity of cytoplasm on the relaxation phenomenon of red blood cell suspensions, we calculate the complex intrinsic viscosity [eta*] = lim(eta* - eta)/eta c of the disperse system of spherical c----0 cells as a function of the frequency, where eta* is the complex viscosity in suspensions, eta the medium viscosity and c the volume concentration of the cells. The cell consists of a viscoelastic membrane and a viscoelastic cytoplasm. The viscoelasticity of the membrane is described by the Voigt model, while the viscoelasticity of the cytoplasmic region is described either by the Maxwell model or by the Voigt model. The interfacial tension is taken into account on both the interfaces of the membrane. The results of [eta*] are compared with the ones in the case in which the cytoplasmic region is purely viscous liquid.  相似文献   

19.
The triply phosphorylated form of human beta-casein comprises about 15% of that fraction and is thus a significant component about midway between the two extremes of zero and five phosphoryls. Its partial specific volume, v, of 0.74 +/- 0.01 and absorbancy, E1% 1 cm, 280 nm, of 6.2 +/- 0.2 are almost identical to the other human beta-caseins. Equilibrium dialysis gave an average of 3.1 +/- 0.4 major Ca2+ binding sites at 37 degrees C with Kdiss = 8.6 x 10(-4) M. Sedimentation and viscosity at low temperatures or in 3.3 M urea suggested a prolate ellipsoidal monomer with 1.4 g H2O/g protein, 10 nm in length and 1.4 nm in width. The concentrated charge of the phosphoryls may be near one end of the ellipsoid, allowing the molecules to align with the flow in the viscometer at low concentration but, due to intermolecular electrostatic interactions, not when concentration is high. This would provide a reason for the heretofore unexplained curvature in the plots of reduced viscosity, eta red, vs beta-casein protein concentration. Self-association increased with temperature. At 37 degrees C in low salt buffer, s20,W was 16 S, which increased to about 33 S as ionic strength, I, was increased to 0.2 and above. At the same time, eta red in low salt buffer decreased from about 22 ml/g at 4 degrees C to a constant value of about 5 ml/g above 23 degrees C. A similar value for eta red at 37 degrees C, which was almost independent of protein concentration, was obtained at I greater than 0.25, giving an extrapolated intrinsic viscosity value of [eta] = 4.0 ml/g. Using this value and assuming a spherical aggregate, calculations suggest a radius of 9 nm with about 48 monomers and 0.86 g H2O/g protein.  相似文献   

20.
Solution properties of tragacanthin (the water-soluble part of gum tragacanth) were studied by gel permeation chromatography (GPC) combined with multi-angle light scattering and viscometry at 25 degrees C. Photon correlation spectroscopy was used to determine the hydrodynamic radius. Ultrasonic degradation was applied to obtain biopolymer fractions of different molecular weights. The dependence of intrinsic viscosity [eta] and radius of gyration (s2)z(1/2) on weight average molecular mass M(w) for this biopolymer were found to be [eta] = 9.077 x 10(-5) M(w)(0.87) (dL g(-1)) and (s2)z(1/2) in the range of M(w) from 1.8 x 10(5) to 1.6 x 10(6). The conformational parameters of tragacanthin were calculated to be 1111 nm for molar mass per unit contour length (M(L)), 26 nm for persistence length (q) and 1.87 ratio of R(g)/R(h). It was found that the Smidsr?d parameter B, the empirical stiffness parameter was 0.013, which is lower than that of several polysaccharides indicating the stiff backbone for tragacanthin. The rheological behavior of aqueous solutions of gum tragacanth and its insoluble and soluble fractions (bassorin and tragacanthin, respectively) were studied. For concentrations equal to 1%, at 25 degrees C and in the absence of salt, bassorin solution showed the highest viscosity and shear thinning behaviour. Power law and Williamson models were used to describe the rheological behaviour of bassorin and tragacanthin, respectively. Oscillatory shear experiments showed a gel like structure for the bassorin but for tragacanthin the oscillatory data were as would be expected for semi-dilute to concentrated solution of entangled, random coil polymers. NaCl changed the steady and oscillatory rheological properties of both fractions and in this way the final viscosity of bassorin was even less than tragacanthin. The calculated activation energy for bassorin and tragacanthin indicated a more rapid decrease in viscosity with temperature for tragacanthin. The plot of eta(sp,0) versus C[eta] revealed that the transition from dilute to semi-dilute regime occurs at C*[eta] = 2.82 for tragacanthin.  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