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 46 毫秒
1.
Pectin was acid extracted from orange albedo by steam injection heating under pressure. Extraction times ranged from 2 to 6 min at a pressure of about 15 psi. Solubilized pectin was characterized by HPSEC with online light scattering and viscosity detection. Molar mass (M), radius of gyration (R(g)), and intrinsic viscosity ([eta]) all decreased with increasing extraction time when heating temperature was 120 degrees C. At heating times of 3 min, M(w) ranged from 4.9 to 4.5 x 10(5), R(gz) was about 44 nm, and [eta](w) ranged from 8.4 to 7.9 dL/g. Chromatography revealed that solubilized pectin distributions were bimodal in nature at 3 min extraction time and trimodal when the extraction time was 6 min. Scaling law exponents obtained for the highest molar mass fractions were consistent with a very compact spherical structure. For the intermediate fraction, scaling law exponents were consistent with a less compact spherical structure comparable to a random coil. In the case of the low molar mass fractions, scaling law exponents were consistent with a structure more asymmetric in shape. These results are consistent with earlier results which indicated that pectin distributions were mixtures of two or more of the following due to disaggregation during extraction: spherical aggregates, hydrogen bonded network structures, and partially or fully disaggregated components of network structures which could include branched structures, rods, segmented rods, and kinked rods.  相似文献   

2.
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   

3.
Nucleohistone solubilized from rabbit thymus nuclei by an endogenous nuclease has in 0.15 M salt an exceptionally low intrinsic viscosity and very high sedimentation velocity. A fully reversible expansion of configuration occurs on lowering ionic strength. When [eta] is plotted against I-1/2 and extrapolated to high I, [eta] = 0 is reached at I = 0.4-1 M and [eta] at I = infinity is negative, contrary to the behavior of DNA and of the great majority of polyelectrolytes, which extrapolate to a positive [eta] at I = infinity. This behavior demands that the configuration of nucleohistone depends not only on electrostatic expansive forces but also on contracting forces which are not electrostatic and do not go to zero in any accessible configuration. Intramolecular hydrophobic bonds might provide such contracting forces. Increasing I above 0.15 M leads to precipitation near 0.3 M and redissolution with dissociation of F1 and expansion in 0.6 M. The expansion is largely but not completely reversed on return to 0.15 M. Much further expansion occurs in I = 1.2 M. Nucleohistone exposed to 1.2 M could not be redissolved in the original medium. Nucleohistone depleted of F1 exhibits a similar expansion as ionic strength is reduced, at higher viscosities throughout. On extrapolation to I = infinity both positive and negative viscosities were observed, on different lots, perhaps reflecting variable extraction of other histones. Circular dichroism spectra are very little affected by ionic strength (0.6 M and lower) or F1 removal, despite tenfold changes in viscosity.  相似文献   

4.
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   

5.
The structure of six human myeloma proteins: IgG1(Bal), IgG2(Klu), IgG3(Bak), IgG3(Het), IgG4(Kov) and IgG4(Pol), was studied in solution using small-angle X-ray scattering and hydrodynamic methods. For IgG1(Bal) and IgG3(Het) the experimental data, including radius of gyration (Rg degree), radii of gyration of the cross-section (Rq1, Rq2), intrinsic viscosity [eta], sedimentation coefficient (S degree 20,w) and molecular mass, were interpreted in terms of structural models based on the Fab and Fc conformations, observed in crystal, by varying the relative positions of the Fab and Fc parts, i.e. their relative angles and distances. The values Rg degree = (6.00 +/- 0.05) nm, S degree 20,w = (6.81 +/- 0.10) S and [eta] = 0.0062 +/- 0.0005 cm3/mg obtained for IgG1(Bal) are compatible with a planar model in which the angle between the Fab arms is about 120 degrees. For IgG3(Het) the following data were obtained: Rg degree = (4.90 +/- 0.05) nm, S degree 20,w = (6.32 +/- 0.01) S and [eta] = (0.0065 +/- 0.0005) cm3/mg. The apparent contradiction between the higher molecular mass and lower Rg degree and S degree 20,w values for IgG3(Het) in comparison to IgG1(Bal) can be resolved by proposing a 'non-planar' (tetrahedral) molecular shape, in which the long hinge peptide is in a folded conformation and the two Fab and Fc parts are in a closely packed arrangement. In this model the angle between the two Fab arms is about 90 degrees, in the average position. The X-ray scattering and hydrodynamic behaviour of the IgG2 and IgG4 types of antibodies appeared to be similar to IgG1(Bal). The parameters of the two IgG3 proteins are similar while they are different to the others.  相似文献   

6.
Physical and chemical properties of LPPS in aqueous solutions were studied. Hydrodynamic characteristics of LPPS depend upon the input concentration and method of solution preparation. Equation parameters for the relationship between [eta] and M omega were determined, and it was found that LPPS behavior in solution can be determined by the equation: [eta] = 5.5.10(-4).M0.57. In aqueous solutions complex macromolecules form compact particles ranging from 8 to 14 nm in size.  相似文献   

7.
Pennisetin, the alcohol soluble storage protein of pearl millet (Pennisetum americanum), was isolated in a homogeneous state. The intrinsic viscosity [n] of this protein was found to be in the range of 16.5-17.7 ml/g in 70% (v/v) aqueous ethanol. The [eta] changed marginally when temperature was increased from 20 to 70 degrees C and also in the presence of 10 mM NaCl. The data indicated that pennisetin was a rigid, rod shaped asymmetric hydrodynamic particle with molecular dimensions in the range of 301 x 14.4 A - 317.7 x 14.2 A. During denaturation with guanidine hydrochloride (Gdn.HCl), the intrinsic viscosity of pennisetin increased from 16 to 25ml/g with a mid point at 3.6 M of the denaturant. The native protein structure was unfolded in 6 M Gdn.HCl as shown by the exposure of aromatic amino acid residues buried in the native state and this transition was found to be reversible. The intrinsic viscosity of pennisetin in 5.9 M Gdn.HCl corresponded to Mr 25,000 which was comparable to that determined by SDS-PAGE.  相似文献   

8.
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   

9.
The effects of temperature and concentration on the viscosity of orange peel pectin solutions were examined at five different temperatures between 20 and 60°C and five concentration levels between 2.5–20 kg/m3. The effects of temperature was described by an Arrhenius-type equation. The activation energy for viscous flow was in the range 19.53–27.16 kJ/mol, depending on the concentration. The effect of concentration was described by two types of equation, power-law and exponential. Equations were derived which describes the combined effects of temperature and concentration on the viscosity for two different models in the range of temperatures and concentrations studied. Orange peel pectin was extracted by using HCl (pH 2.5, 90°C, 90 min) ammonium oxalate (0.25%, pH 3.5, 75°C, 90 min) and EDTA (0.5%, 90°C, 90 min) extraction procedures. The best result was obtained with ammonium oxalate extraction in which the pectin content of the final product was 30.12%, although the efficiency among the procedures varied.The average molecular weight was measured by light scattering technique. Magnitudes of intrinsic viscosity and molecular weight of pectins obtained by extraction with HCl, ammonium oxalate and EDTA were 0.262, 0.281, 0.309 m3/kg and 84 500, 91 400, 102 800 kg/kgmol, respectively. The molecular weight dependence of the intrinsic viscosity of the orange peel pectin solutions was expressed by Mark–Houwink–Sakurada equation. The data were fitted to equation as ηi=2.34×10−5(Mw,ave)0.8224 which helps to evaluate the average molecular weight of pectin solutions from orange peel with a knowledge of their intrinsic viscosity.  相似文献   

10.
The white part of citrus peel, the albedo, has a special role in water relations of both fruit and leaves from early on in fruit development. In times of drought, this tissue acts as a water reservoir for juice sacs, seeds and leaves. When water was injected into the albedo, free water was undetectable using magnetic resonance imaging. Microscopy showed tightly packed cells with little intercellular space, and thick cell walls. Cell wall material comprised 21% of the fresh albedo weight, and contained 26.1% galacturonic acid, the main constituent of pectin. From this, we postulated that pectin of the cell wall was responsible for the high water-binding capacity of the immature lemon albedo. Cell wall material was extracted using mild procedures that keep polymers intact, and four pectic fractions were recovered. Of these fractions, the SDS and chelator-soluble fractions showed viscosities ten and twenty times higher than laboratory-grade citrus pectin or the other albedo-derived pectins. The yield of these two pectins represented 28% of the cell walls and 62% of the galacturonic acid content of immature lemon albedo. We concluded that, from viscosity and abundance, these types of pectin account for the high water-binding capacity of this tissue. Compositional analyses showed that the two highly viscous pectic fractions differ in galacturonic acid content, degree of branching and length of side chains from the less viscous albedo-derived pectins. The most striking feature of these highly viscous pectins, however, was their high molecular weight distribution compared to the other pectic fractions.  相似文献   

11.
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   

12.
The present study demonstrates for the first time the solubilization of peripheral-type benzodiazepine binding sites (PBS) from cat cerebral cortex. Of all detergents tested [digitonin, 3-[(3-cholamidopropyl)dimethylammonio]-1-propane sulfonate (CHAPS), Tween 20, deoxycholate, and Triton X-100] in the presence of NaCl, the best solubilization (15% of initial activity) was obtained using 0.5% of the zwitterionic detergent CHAPS plus 2 M NaCl. Specific binding of [3H]PK 11195 to membrane-bound and solubilized PBS was saturable, yielding equilibrium dissociation constants (KD) of 1.3 +/- 0.2 and 1.9 +/- 0.3 nM, respectively, and maximal numbers of binding sites of 1,435 +/- 150 and 980 +/- 126 fmol/mg protein, respectively. The KD value of PK 11195 binding to solubilized PBS obtained from experimental kinetic analysis was 0.95 +/- 0.09 nM. The relative potencies of various compounds (PK 11195, Ro 5-4864, diazepam, flunitrazepam, clonazepam, methyl-beta-carboline-3-carboxylate, and Ro 15-1788) in displacing [3H]PK 11195 specific binding from membrane-bound and solubilized PBS were similar. Most of the solubilized binding activity was destroyed by heating at 60 degrees C for 30 min or by treatment with 2 M guanidinium chloride, which indicates the presence of a protein-binding site in the solubilized preparation. Over 85% of the solubilized binding activity was retained after 1 week at 4 degrees C, which will enable future application of purification procedures without major concern for stability of the material.  相似文献   

13.
The specific binding protein for substance P (SP) was solubilized in an active form from the crude mitochondrial (P2) fraction of bovine brainstem. After incubation with 3-[(3-cholamidopropyl)dimethylammonio]-1-propane sulfonate (CHAPS) and 0.1 M NaCl at 0 degrees C for 30 min, the SP binding to the supernatant fraction (100,000 g, 60 min) was determined by the glass fiber filtration method reported by Bruns et al. (1983). The specific [3H]SP binding to the solubilized fraction was highly specific for SP and was displaced by nanomolar concentrations of SP and physalaemin, but only by micromolar concentrations of eledoisin. In addition, the binding was inhibited by GTP (approximately 40% of the specific binding decreased by 10 microM GTP) in both preparations. These results were virtually identical to those of P2 membrane preparations and suggested that this high-affinity SP binding site belongs to the SP-P type. Scatchard analyses of SP binding to the solubilized fraction revealed a single saturable component with a Bmax of 22.0 +/- 5.10 fmol/mg protein and a KD of 0.79 nM, and these values are almost the same as those obtained in the P2 fraction (Bmax = 31.3 +/- 3.56 fmol/mg protein, KD = 0.82 nM). Gel filtration analysis showed that the detergent-SP binding protein complex has two calculated molecular weights of greater than 1,000,000 and 55,000-60,000 (a corresponding Stokes radius of 35.5 nm).  相似文献   

14.
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   

15.
The specific binding protein for prostaglandin (PG) E2 was solubilized in an active form from the crude mitochondrial (P2) fraction of porcine cerebral cortex. After incubation with 3-[(3-cholamidopropyl)dimethylammonio]-1-propane sulfonate (CHAPS) at 4 degree C for 30 min, the PGE2 binding to the supernatant fraction (103,000 g, 60 min) was determined by the polyethylene glycol method. The maximum yield (approximately 30% of the binding activity to the P2 fraction) was obtained with 10 mM CHAPS. The specific [3H]PGE2 binding to the solubilized fraction was time-dependent and the equilibrium was reached at around 60 min at 37 degrees C. By dilution of the reaction mixture, the binding site-[3H]PGE2 complex formed after 5-min incubation slowly dissociated, whereas that formed after 60-min incubation did not dissociate to a significant extent. The binding was highly specific for PGE2 and inhibited by unlabeled PGs in the following order: PGE2 greater than PGE1 much greater than PGF2 alpha greater than PGE2 methyl ester greater than PGA2 greater than 13,14-dihydro-15-keto-PGE2 greater than PGD2. Scatchard analyses of the solubilized fraction suggested the presence of high- and low-affinity sites. Heat treatment and preincubation with trypsin or proteinase K markedly reduced the binding. The binding activity was eluted in a single peak both from gel filtration and from ion-exchange columns using HPLC. These results suggest that a specific protein solubilized may be responsible for the binding site.  相似文献   

16.
With raising temperature in the domain of 20 to 60 degrees C, the intrinsic viscosity [eta] for pullulan-tricarbanilate PTC and amylose-tricarbanilate ATC in solutions was found to decrease, indicating that they exhibited thermal-induced conformational transition from expanded form to compact form. The persistence length P(l) of the chains, evaluated with small-angle X-ray scattering, has also decreased as the temperature is raised and, moreover, it significantly depended on the solvents employed, where as P(l) of pullulan, having no carbanilate groups, has exhibited neither temperature- nor solvent-dependence. The temperature dependence of [eta] for PTC and ATC was well elucidated in terms of the temperature-dependent P(l) with the wormlike chain model. From these results, it is suggested that intramolecular hydrogen bonds would be formed between carbanilate groups neighboring along the backbone chain, but they are gradually and cooperatively collapsed as the temperature is raised, inducing the conformational transition.  相似文献   

17.
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   

18.
ABSTRACT: BACKGROUND: The bacterium Acetobacter sp. CCTCC M209061 is a promising whole-cell biocatalyst with exclusive anti-Prelog stereoselectivity for the reduction of prochiral ketones that can be used to make valuable chiral alcohols such as (R)-4-(trimethylsilyl)-3-butyn-2-ol. Although it has promising catalytic properties, its stability and reusability are relatively poor compared to other biocatalysts. Hence, we explored various materials for immobilizing the active cells, in order to improve the operational stability of biocatalyst. RESULTS: It was found that Ca-alginate give the best immobilized biocatalyst, which was then coated with chitosan to further improve its mechanical strength and swelling-resistance properties. Conditions were optimized for formation of reusable immobilized beads which can be used for repeated batch asymmetric reduction of 4[prime]-chloroacetophenone. The optimized immobilized biocatalyst was very promising, with a specific activity of 85% that of the free-cell biocatalyst (34.66 mumol/min/g dw of cells for immobilized catalyst vs 40.54 mumol/min/g for free cells in the asymmetric reduction of 4[prime]-chloroacetophenone). The immobilized cells showed better thermal stability, pH stability, solvent tolerance and storability compared with free cells. After 25 cycles reaction, the immobilized beads still retained >50% catalytic activity, which was 3.5 times higher than degree of retention of activity by free cells reused in a similar way. The cells could be recultured in the beads to regain full activity and perform a further 25 cycles of the reduction reaction. The external mass transfer resistances were negligible as deduced from Damkohler modulus Da < <1, and internal mass transfer restriction affected the reduction action but was not the principal rate-controlling step according to effectiveness factors eta < 1 and Thiele modulus 0.3<[empty set] <1. CONCLUSIONS: Ca-alginate coated with chitosan is a highly effective material for immobilization of Acetobacter sp. CCTCC M209061 cells for repeated use in the asymmetric reduction of ketones. Only a small cost in terms of the slightly lower catalytic activity compared to free cells could give highly practicable immobilized biocatalyst.  相似文献   

19.
This work evaluates linoleic acid peroxidation reactions initiated by Fe3+-reducing compounds recovered from Eucalyptus grandis, biotreated with the biopulping fungus Ceriporiopsis subvermispora. The aqueous extracts from biotreated wood had the ability to reduce Fe3+ ions from freshly prepared solutions. The compounds responsible for the Fe3+-reducing activity corresponded to UV-absorbing substances with apparent molar masses from 3 kDa to 5 kDa. Linoleic acid peroxidation reactions conducted in the presence of Fe3+ ions and the Fe3+-reducing compounds showed that the rate of O2 consumption during peroxidation was proportional to the Fe3+-reducing activity present in each extract obtained from biotreated wood. This peroxidation reaction was coupled with in-vitro treatment of ball-milled E. grandis wood. Ultraviolet data showed that the reaction system released lignin fragments from the milled wood. Size exclusion chromatography data indicated that the solubilized material contained a minor fraction representing high-molar-mass molecules excluded by the column and a main low-molar-mass peak. Overall evaluation of the data suggested that the Fe3+-reducing compounds formed during wood biodegradation by C. subvermispora can mediate lignin degradation through linoleic acid peroxidation.  相似文献   

20.
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   

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