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.
Following the work carried out on guar gum in our first paper of a series, the "pressure cell" solubilization method was applied to two other less highly substituted galactomannans: locust bean gum (LBG) and tara gum. True molecular solution of the polymers was achieved using appropriate temperature, time, and pressure regimes. The technique of capillary viscometry was used to determine the intrinsic viscosity [eta] of the "pressure cell" treated and untreated samples. Molecular weight (M(w)) and radius of gyration (R(g)) were determined by light scattering. The data obtained for LBG and tara gum were compared statistically with reliable data found for guar gum in the literature. The variation in [eta] with M(w) followed the Mark-Houwink-Sakurada relationship, giving the exponent alpha = 0.74 +/- 0.01 for galactomannans consistent with random coil behavior. The characteristic ratio, C(infinity), and the chain persistence length, L(p), were both calculated for LBG and tara gum using the Burchard-Stockmayer-Fixman (BSF) method which is appropriate for flexible to semiflexible chains. A general value of 9 < C(infinity) < 16 and 3 < L(p) < 5 nm can now be estimated with statistical confidence for all galactomannans. According to our statistical analysis, the chain persistence length was found to be insensitive to the degree of galactose substitution.  相似文献   

2.
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   

3.
The influence of the galactomannan characteristic ratios (M/G) on the temperature of gelation (Tg) and the gel strength of mixtures of galactomannan with xanthan is reported. Two galactomannans were investigated: one highly substituted from the seeds of Mimosa scabrella (M/G = 11), and the other, less substituted, from the endosperm of Schizolobium parahybae, with (M/G = 30) [Ganter JLMS, Zawadzki-Baggio SF, Leitner SC, Sierakowski MR, Reicher F. J Carbohydr Chem 1993;12:753]. The xanthan:galactomannan systems (4:2 g l(-1), in 5 mM NaCl) showed a temperature of gel formation (Tg) of 24 degrees C for that of S. parahybae [Bresolin TMB, Milas M, Rinaudo M and Ganter JLMS. Int J Biol Macromol 1998;23:263] and 20 degrees C for the galactomannan of M. scabrella, determined by viscoelastic measurements and microcalorimetry. A Tg of 40-50 degrees C was found by Shatwell et al. [Shatwell KP, Sutherland IW, Ross-Murphy SB, Dea ICM. Carbohydr Polym 1991;14:29] for locust bean gum-LBG (M/G = 43). Lundin and Hermansson [Lundin L, Hermansson AM. Carbohydr Polym 1995;26:129] reported a difference of 13 degrees C for Tg of two LBG samples with M/G = 3 (40 degrees C) and 5 (53 degrees C), in mixtures with xanthan. It appears that the more substituted galactomannans have lower temperatures of gelation in the presence of xanthan. The mechanism of gelation depends also on the M/G ratio. For the lower values it involves only disordered xanthan chains in contrast to M/G ratios higher than 3. In addition, the presence of the galactomannan from M. scabrella increased slightly the temperature of the conformational change (Tm) of xanthan probably due to the ionic strength contribution of proteins (3.9%) present in the galactomannan. On the other hand, the galactomannans from S. parahybae, with 1.5% of proteins and M. scabrella, with 2.4% of protein, did not show this effect, the Tm of xanthan alone or in a mixture being practically unchanged.  相似文献   

4.
5.
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   

6.
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   

7.
H X Zhou 《Biophysical journal》1995,69(6):2286-2297
A general method for calculating translational friction and intrinsic viscosity is developed through exploiting relations between hydrodynamics and electrostatics. An approximate relation xi = 6 pi eta 0C between the translational friction coefficient xi of a particle (eta 0: solvent viscosity) and its capacitance C was derived previously. This involved orientationally preaveraging the Oseen tensor, but the result was found to be very accurate. Based on preaveraging, we find that the intrinsic viscosity [eta] of a particle can be estimated from its polarizability alpha through [eta] = 3/4 alpha + 1/4 Vp, where Vp is the volume of the particle. Both the capacitance and the polarizability can be obtained in a single calculation using the boundary-element technique. An efficient approach is thus found for estimating [eta], a quantity that is very useful in practice because of its sensitivity to particle shape but is notoriously difficult to calculate. Illustrative calculations on ellipsoids, cylinders, and dumbbells demonstrate both the accuracy of the approximate relations and the efficiency of the present method.  相似文献   

8.
Nine hyaluronan (HA) samples were fractionated by size-exclusion chromatography, and molar mass (M), radius of gyration (Rg), and intrinsic viscosity ([eta]) were measured in 0.15 M NaCl at 37 degrees C by on-line multiangle light scattering and viscometer detectors. Using such method, we investigated the Rg and [eta] molar mass dependence for HA over a very wide range of molar masses: M ranging from 4 x 10(4) to 5.5 x 10(6) g/mol. The Rg and the [eta] molar mass dependence found for HA showed a meaningful difference. The Rg = f(M) power law was substantially linear in the whole range of molar masses explored with a constant slope of 0.6. In contrast, the [eta] = f(M) power law (Mark-Houwink-Sakurada plot) showed a marked curve shape, and a linear regression over the whole range of molar masses does not make sense. Also the persistence length (stiffness) for HA was estimated. The persistence length derived by using both the Odijk's model (7.5 nm from Rg vs M data) and the Bohdanecky's plot (6.8 nm from [eta] vs M data) were quite similar. These persistence length values are congruent with a semistiff conformation of HA macromolecules.  相似文献   

9.
A water soluble galactomannan isolated from Leucaena leucocephala seeds gave an intrinsic viscosity of 3.5dl/g and viscosity average molecular mass, M(v), of 6.98×10(5)g/mol. This was in reasonably good agreement with the value of the weight average molecular mass, M(w), of 5.44±0.20×10(5)g/mol determined by GPC-MALLS coupled to RI. The onset of polymer coil overlap occurred at c*[η] of 2.1, with slope of 3.0 above and 1.3 below the point of polymer coil overlap. The shear viscosity of the polysaccharide was temperature dependent and decreased with increasing temperature. The activation energy for viscous flow of 3.0% polysaccharide concentration obtained by Arrhenius plot of zero shear viscosity as a function of temperature was 26.4kJ/mol. Both the storage modulus (G') and loss modulus (G″) showed strong dependence on frequency indicating the presence of entangled coils. The Cox-Merz plot gave close superimposition of the complex and shear viscosities.  相似文献   

10.
The rotational dynamics of TEMPAMINE can be used to study directly the intracellular environment. The extracellular signal from TEMPAMINE is broadened away by the use of potassium ferricyanide which does not enter the cell. The EPR signal which results when 1 mM TEMPAMINE, 120 mM ferricyanide, and erythrocytes are mixed together arises from TEMPAMINE only in the intracellular aqueous space. The relative viscosity measured by the motion of TEMPAMINE in various control environments is: water at 37 degrees C = 1; human plasma at 37 degrees C = 1.1; internal aqueous environment of washed erythrocytes or whole blood at 37 degrees C = 4.92 +/- 0.32. Erythrocytes can be fractionated by density. In sickle-cell anemia (SS), the percentage of cells we find with density greater than 1.128 g/ml is 15-40%, in normals (AA) and sickle trait (AS) 1%. By direct spin-label measurements with TEMPAMINE we show, for the first time, that the relative internal viscosity (eta mu) of these dense erythrocytes is markedly elevated and density-dependent. Our results show that (1) eta mu increases with increasing cell density; (2) eta mu obtained from sickle cells is higher than eta mu obtained from normal cells at a given density, and this effect is greater at 37 degrees C than at 20 degrees C; (3) eta mu is proportional to MCHC, but eta mu in erythrocytes is higher than eta mu obtained from in vitro preparations of hemoglobin S at equivalent concentrations. We conclude that the relative internal viscosity of erythrocytes is affected by three factors: the state of cell hydration, the amount of hemoglobin polymer present, and the potential interactions of the cell membrane with intracellular hemoglobin.  相似文献   

11.
The melt curve and the liquid-state transport properties shear viscosity, self-diffusion coefficient and thermal conductivity of 1,3,5-triamino-2,4,6-trinitrobenzene (TATB) were predicted using all-atom molecular dynamics simulations. The TATB melt curve was obtained using solid–liquid coexistence simulations and is in good accord with the Simon–Glatzel equation. The temperature dependencies of the shear viscosity and self-diffusion coefficient are predicted to obey Arrhenius behaviour for pressures up to P = 20 kbar. The thermal conductivity has a linear temperature dependence for P < 15 kbar and a linear density (ρ) dependence for ρ > 1200 kg m?3. At similar densities the shear viscosity of liquid TATB is close to the predictions for liquid nitromethane [58] but lower than the predictions for liquid HMX [24] and RDX [59]. The self-diffusion coefficient for TATB is predicted to be higher than predictions for nitromethane, HMX and RDX at similar densities. The conductivity of TATB is ≈20% greater than the conductivity of liquid HMX at a given density.  相似文献   

12.
The viscosity in the low shear rate Newtonian domain of three biopolymers, locust bean gum, guar gum and xanthan gum was studied as a function of temperature and of polymer concentration in various aqueous solvents. The intrinsic viscosities [η]o of both galactomannans are not modified in the presence of 10 or 40% sucrose. In this case, a master curve relating the Newtonian specific viscosity (ηsp)o, to the reduced concentration c[η]o is obtained and allows (in good agreement with theoretical conjectures), two critical concentrations C* and C** to be defined, from which the value of the expansion coefficient may be estimated. For xanthan, as expected for a polyelectrolyte, [η]o depends strongly on salt concentration and on added sucrose and the results did not obey the above-mentioned master curve. However, it is shown that (ηsp)o depends only on xanthan concentration whenC > C**, and then it is assumed that chain dimensions have attained their unperturbed values whatever the solvent. Considering that both types of chains, random coils (galactomannans) and semi-rigid (xanthan) should give the same (ηsp)o-C[η]o master curve for C > C** when [η]o is replaced by its unperturbed counterpart [η]θ, a method for estimating [η]θ for the xanthan sample is proposed. In conclusion, the numerous exceptions to the widely accepted (ηsp)o vs C[η]o “universal” behaviour are mainly ascribed to significant differences in expansion coefficient values which depend on both the polymer and the solvent.  相似文献   

13.
Red blood cell orientation in orbit C = 0.   总被引:4,自引:0,他引:4       下载免费PDF全文
M Bitbol 《Biophysical journal》1986,49(5):1055-1068
Two modes of behavior of single human red cells in a shear field have been described. It is known that in low viscosity media and at shear rates less than 20 s-1, the cells rotate with a periodically varying angular velocity, in accord with the theory of Jeffery (1922) for oblate spheroids. In media of viscosity greater than approximately 5 mPa s and sufficiently high shear rates, the cells align themselves at a constant angle to the direction of flow with the membrane undergoing tank-tread motion. Also, in low viscosity media, as the shear rate is increased, more and more cells lie in the plane of shear, undergoing spin with their axes of symmetry aligned with the vorticity axis of the shear field in an orbit "C = 0" (Goldsmith and Marlow, 1972). We have explored this latter phenomenon using two experimental methods. First, the erythrocytes were observed in the rheoscope and their diameters measured. Forward light scattering patterns were correlated with the red cell orientation mode. Light flux variations after flow onset or stop were measured, and the characteristic times of erythrocyte orientation and disorientation were assessed. The characteristic time of erythrocyte orientation in Orbit C = 0 is proportional to the inverse of the shear rate. The corresponding coefficient of proportionality depends on the suspending medium viscosity eta o. The disorientation time tau D, after flow has been stopped, is such that the ratio tau D/eta o is independent of the initial applied shear stress. However, tau D is much shorter than one would expect if pure Brownian motion were involved. The proportion of erythrocytes in orbit C = 0 was also measured. It was found that this proportion is a function of both the shear rate and eta o. At low values of eta o, the proportion increases with increasing shear rate and then reaches a plateau. For higher values of eta o (5 to 10 mPa s), the proportion of RBC in orbit C = 0 is a decreasing function of the shear stress. A critical transition between orbit C = 0 and parallel alignment was observed at high values of eta o, when the shear stress is on the order of 1 N/m2. Finally, the effect of altering membrane viscoelastic properties (by heat or diamide treatment) was tested. The proportion of oriented cells is a steep decreasing function of red cell rigidity.  相似文献   

14.
The composition and structure of a galactomannan from seeds of Astragalus lehmannianus, an endemic legume species, is reported for the first time. The purified galactomannan (yield, 4.8%) contained 55% D-mannose and 45% D-galactose and had a molecular weight of 997.03 kDa. Its aqueous solutions were optically active and highly viscous (the specific rotation, [alpha]D, equaled +81.3 degrees; the characteristic viscosity, [eta], 868.4 ml/g). Chemical, chromatographic, and spectral (IR and 13C-NMR spectroscopy) methods were used to demonstrate that the main chain of the molecule is formed by residues of 1,4-beta-D-mannopyranose, 78% of which are substituted at position 6 with single alpha-D-galactopyranose. The distribution of galactose along the chain was calculated from NMR spectra: frequencies of occurrence, per pair of neighboring mannose units, of (1) two substituents, (2) one substituent, and (3) no substituents were 65.3, 31.5, and 3.2%, respectively. The specific rotation of galactomannans was shown to correlate with their content of galactose.  相似文献   

15.
The intrinsic viscosity ([eta]) and the molecular weight (M) by sedimentation equilibrium were determined for hyaluronic acids of low (M=104--7.2X10(4)) and high (M=3.1X10(5)--1.5X10(6)) molecular weights. Double logarithmic plot of [eta] against M gave different lines for the two groups. The relationship between [eta] and M was [eta]=3.0X10(6)XM1,20 for the former and [eta]=5.7X10(-4)XM0.46 for the latter group. The molecular weight at the point of intersection of the two lines was about 1.5X10(5). The rheological behavior of the hyaluronic acids below M=2.1X10(4), for which the value of reduced viscosity was independent of concentration, was different from that of the hyaluronic acids above M=5.1X10(4), for which the value of reduced viscosity increased with concentration.  相似文献   

16.
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   

17.
We report the rheological behaviour of a succinoglycan polysaccharide in dilute and semi-dilute solutions as a function of temperature, ionic strength and the nature of counterion. We have examined the viscosity dependence as a function of molecular weight using samples obtained by ultrasonication. We have also prepared samples lacking succinate substitutes and compared their behaviour with that of the native polymer. In both cases, we observed that, after heating a polymer solution for the first time above the conformational transition temperature, a different ordered state was obtained on cooling. This state had a lower molecular weight and intrinsic viscosity but identical chemical structure and local properties. A role for the side chain in the stabilization of breaks in the backbone is suggested. Nevertheless, a unique curve is obtained for the specific viscosity as a function of the overlap parameter c[eta] for different polymer concentrations of both the native and heated forms. However, different curves are obtained for normal and succinate-free polymers, and the succinate-free polymer is characterized by a lower Huggins constant.  相似文献   

18.
Shear deformation of young and old human red blood cells was examined over a range of shear stresses and suspending phase viscosities (eta o) using a cone-plate Rheoscope. The internal viscosities (eta i) of these cell types differ, and further changes in internal viscosity were induced by alteration of suspension osmolality and hence cell volume. For low suspending viscosities (0.0555 or 0.111 P) old cells tended to tumble in shear flow, whereas young cells achieved stable orientation and deformed. Changes in osmolality, at these external viscosities, altered the percentage of cells deforming, and for each cell type threshold osmolalities (Osm-50) were determined where 50% of cells deformed. The threshold osmolalities were higher for younger cells than for older cells, but the internal viscosities of the two cell types were similar at their respective Osm-50. Threshold osmolalities were also higher for the higher external viscosity, but the ratio of internal to external viscosities (i.e., eta i/eta o) was nearly constant for both external viscosities. Deformation of stably oriented cells increased with increasing shear stress and approached a value limited by cell surface area and volume. For isotonic media, over a wide range of external viscosities and shear stresses, deformation was greater for younger cells than for older cells. However, deformation vs. shear stress data for the two cell types became nearly coincident if young cells were osmotically shrunk to have their internal viscosity close to that for old cells. Increases in external viscosity, at constant shear stress, caused greater deformation for all cells. This effect of external viscosity was not equal for young and old cells; the ratio of old/young cell deformation increased with increasing eta o. However, if deformation was plotted as a function of the ratio lambda = eta i/eta o, at constant shear stress, young and old cell data followed similar paths. Thus the ratio lambda is a major determinant of cell deformation as well as a critical factor affecting stable orientation in shear flow.  相似文献   

19.
Locust bean gum extracted from two carob flours from eastern and western Mediterranean sources were fractionated on the basis of their solubility in water. Weight-average molecular weights determined by sedimentation equilibrium were about 300 000 for both the hot water and cold water soluble fractions, whereas a commercial sample of guar gum had a molecular weight of 700 000. Their values were lower than would be predicted from Mark-Houwink relationships where molecular weights were originally determined by light scattering.

The hot water soluble fraction from the eastern Mediterranean flour showed unexpected rheological behaviour. It had an extremely high Huggins' constant and a different relationship between the coil overlap parameter and the zero shear rate viscosity compared with previously reported results for galactomannans. Both effects may be explained by the anomalously low intrinsic viscosity of this fraction when determined by a Huggins' extrapolation. The use of the Kraemer extrapolation gave significantly higher intrinsic viscosities for this particular sample. Gels formed from the two hot water soluble fractions with κ-carrageenan had similar rheological properties.  相似文献   


20.
The viscosity of ovalbumin aqueous solutions was studied as a function of temperature and of protein concentration. Viscosity-temperature dependence was discussed on the basis of the modified Arrhenius formula at temperatures ranging from 5 to 55 degrees C. The activation energy of viscous flow for hydrated and unhydrated ovalbumin was calculated. Viscosity-concentration dependence, in turn, was discussed on the basis of Mooney equation. It has been shown that the shape parameter S decreases with increasing temperature, and self-crowding factor K does not depend on temperature. At low concentration limit the numerical values of the intrinsic viscosity and of Huggins coefficient were calculated. A master curve relating the specific viscosity etasp to the reduced concentration c[eta], over the whole range of temperature, was obtained and the three ranges of concentrations: diluted, semi-diluted and concentrated, are discussed. It has been proved that the Mark-Houvink-Kuhn-Sakurada (MHKS) exponent for ovalbumin does not depend on temperature.  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