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.
The technique of quasielastic light scattering (QELS) in the presence of a sinusoidal electric field (QELS-SEF) was used to examine the diffusion of poly(lysine) in the extraordinary phase, which is defined by the anomalously small value in the apparent diffusion coefficient (Dapp, obtained by QELS methods) as the added salt concentration is lowered below 10 mM. Spectra were obtained for high-molecular-weight poly(lysine) (800,000) in 0.5 mM KCl, using a driving frequency of 90 Hz. It was observed that the linewidth, hence apparent diffusion coefficient (DSF), increased with electric field strength (E) over the range 6 < E < 30 V/cm. As might be anticipated, the asymptotic limit, DSF(E = 0), was almost numerically equivalent to Dapp at this KCl concentration. As E approached the upper limit value of 31 V/cm, the value of DSF approached the high salt (500 mM KCl) value of Dapp. These observations are discussed in terms of small-ion dispersion effects.  相似文献   

2.
We show that the persistence length a of DNA, derived from total intensity laser light scattering of linear Col E1 DNA and corrected for excluded-volume effects, varies from about 68 nm in 0.005M NaCl to about 40 nm in 0.2M NaCl, leveling off to a constant value (about 27 nm) at high NaCl (1–4M) concentration. These observations do not agree with current views on the effect of electrostatic charge and ionic conditions on DNA dimensions. The apparent diffusion constant Dapp, determined from laser light scattering autocorrelation as a function of scattering vector q, at NaCl concentrations 0.005–4M, correctly yields the translational diffusion coefficient Dt at low values of q and scales with molecular dimensions rather than segment length at high values of q; thus, Dapp/Dt yields a universal curve when plotted against q2R, where Rg is the radius of the gyration. The sedimentation coefficients s at 0.1 and 0.2M NaCl concentration closely agree with the well-tested empirical relations, and a combination of s, Dt, and the appropriate density increments yield correct molar masses over the whole salt concentration range. Approximate constancy of DtRg indicates limited draining in translational flow. We present some observations and thoughts on the regimes in which a dependence of the correlation decay times on q3 rather than q2 applies. We conclude that quasielastic laser light scattering discloses little information about dynamics of internal motion of DNA chains.  相似文献   

3.
In this study, Debaryomyces nepalensis NCYC 3413 isolated from rotten apple was studied for its halotolerance and its growth was compared with that of Saccharomyces cerevisiae in high salt medium. The specific growth rate of D. nepalensis was not affected by KCl even up to a concentration of 1 M, whereas NaCl and LiCl affected the growth of D. nepalensis. Among all tested salts, LiCl showed maximum inhibition on growth. At all conditions, halotolerance of D. nepalensis was much higher than that of S. cerevisiae. D. nepalensis showed maximum viability (80–100%) when grown in KCl, which was higher than with NaCl and LiCl. Pectinase production by D. nepalensis was noted at all high salt concentrations, namely, 2 M NaCl, 2 M KCl, and 0.5 M LiCl, and the maximum specific activity was observed when the strain was grown in 2 M NaCl.  相似文献   

4.
S. Watanabe  T. Saito 《Biopolymers》1987,26(5):625-632
The effect of LiCl, NaCl, and CsCl as univalent salts, and of CaCl2, ZnCl2, and MgCl2 as divalent salts, on the α and antiparallel β-sheet, and random conformations of poly(L-lysine) (PLL), in water at room temperature were examined by means of CD and compared quantitatively on the basis of elliptical strength at the maximal peak. Changes in the α-helical and antiparallel β-sheet helical conformations of PLL were markedly dependent on the salt concentrations of LiCl, NaCl, and CsCl, which induced decreases in negative intensity in that order. The CD spectrum of the random conformation, the most disordered form, displayed positive cotton effect in concentrations of these salts up to 3.0M and a negative peak in concentrations of 6.0M. The effect of these salts on the random conformation of PLL was stronger than that on the α- and β-conformations in higher concentrations. The CD spectrum of the random conformation in the presence of CaCl2, ZnCl2, and MgCl2, on the other hand, showed negative cotton effect in salt concentrations as low as 3.0M. It was impossible, however, to measure the effect on α- and β-conformations of ZnCl2 and MgCl2 above concentrations of 10 mM because of a solubility problem with salts in alkaline solution.  相似文献   

5.
The helical stability of a variety of DNA samples, ranging in base composition from 0 to 72 mole-% GC, has been studied by heat denaturation at neutral pH in increasing concentrations of LiCl, NaCl, KCl, CsCl, Li2SO4, and K2SO4. The variation of melting temperature with average base composition, dTm/dXGC, was found to decrease drastically in the concentrated salt media, e.g., from 41°C in 0.006M LiCl to 29°C in 3.2M LiCl, and from 39°C in 0.003M Li2SO4 to 18°C in 1.6M Li2SO4. At the same time, the thermal transition is much more cooperative in the concentrated salt solutions than at low ionic strength. Indeed, at limiting salt concentrations, the transition breadth seems to reach a minimum value irrespective of the compositional heterogeneity of the DNA samples. Attempts to correlate the observed decrease of dTm/dXGC with predicted changes in the enthalpy of melting, deduced from a simple theoretical treatment, experimental data on the binding of counterions and water to DNA, and experimental data on thermal denaturation, were unsuccessful. However, the strongly reduced composition dependence of the melting temperature can be understood in terms of a destabilizing effect of the concentrated salt media on GC-base pairs. It is suggested, though not proven, that the destabilization involves the displacement of water molecules from the DNA helix.  相似文献   

6.
To understand the root function in salt tolerance, radial salt and water transport were studied using reed plants growing in brackish habitat water with an osmotic pressure (πM) of 0.63 MPa. Roots bathed in this medium exuded a xylem sap with NaCl as the major osmolyte and did so even at higher salt concentration (πM up to 1.3 MPa). Exudation was stopped after a small increase of πM (0.26 MPa) using polyethylene glycol 600 as osmolyte. The endodermis of fine lateral roots was found to be the main barrier to radial solute diffusion on an apoplastic path. Apoplastic salt transfer was proven by rapid replacement of stelar Na+ by Li+ in an isomolar LiCl medium. Water fluxes did not exert a true solvent drag on NaCl. Xylem sap concentrations of NaCl in basal internodes of transpiring culms were more than five times higher than in medial and upper ones. It was concluded that the radial NaCl flux was mainly diffusion through the apoplast, and radial water transport, because of the resistance of the cell wall matrix to convective mass flow, was confined to the symplast. Radial salt permeation in roots reduced the water stress exerted by the brackish medium.  相似文献   

7.
N Murai  S Sugai 《Biopolymers》1974,13(6):1195-1203
The conformational changes of poly-Nε-glutaryl-L -lysine (PGL) and poly-Nε-succinyl-L -lysine (PSL) in various salt solutions were studied by use of ORD and potentiometric titration measurements. The addition of alkali metal salts to the fully ionized PGL or PSL solution caused helix formation. The helical content of the polymers increases in the following sequences: at salt concentration 0–2 M, CsCl < KCl < LiCl < NaCl; and at 2–3 M, LiCl < CsCl < KCl ~ NaCl. The preferential binding of the solvent components with various alkali metal salts of PGL or PSL was measured in LiCl, NaCl, and KCl solutions by means of equilibrium dialysis and differential refractometry. It was found that with increasing salt concentration, the polymers were preferentially hydrated in NaCl and KCl soultions; however the salt was preferentially bound to the polymers in LiCl solution. Such preferential binding was suggested to be closely related to conformational change. The addition of CaCl2 to polymer solutions led to the stabilization of the helical structure of PGL or PSL.  相似文献   

8.
Quasielastic light scattering and electrophoretic light scattering experiments were performed on chicken erythrocyte polynucleosome solutions at various temperatures and ionic strengths. The apparent diffusion coefficient, Dapp, was found to depend on the scattering vector K. In general, Dapp can be described as a damped oscillatory function of K in the ionic strength range of 10 to 60 mM and over the temperature range of 10 to 40°C. Electrophoretic light scattering studies on total digest chromatin samples indicate the apparent charge on the polynucleosomes increases as the ionic strength is lowered from 10 to 1 mM. These data are interpreted in terms of fluctuations in the surface charge distribution of the polyion and subsequent inducement of an asymmetric distribution of small ions about the polyion. These fluctuation components lead to the formation of “clusters” of polyions.  相似文献   

9.
Abstract: [3H]Ryanodine binding studies of ryanodine receptors in brain membrane preparations typically require the presence of high salt concentrations in assay incubations to yield optimal levels of binding. Here, radioligand binding measurements on rat cerebral cortical tissues were conducted under high (1.0 M KCI) and low (200 mM KCI) salt buffer conditions to determine the effects of ionic strength on receptor binding properties as well as on modulation of ligand binding by Ca2+, Mg2+, β,γ-methylene-adenosine 5′-triphosphate (AMP-PCP), and caffeine. In 1.0 M KCI buffer, labeled titration/equilibrium analyses yielded two classes of binding sites with apparent KD (nM) and Bmax (fmol/mg of protein) values of 2.4 and 34, respectively, for the high-affinity site and 19.9 and 157, respectively, for the low-affinity site. Unlabeled titration/equilibrium measurements gave a single high-affinity site with a KD value of 1.9 nM and a Bmax value of 95 fmol/mg of protein. The apparent KD value derived from association and dissociation studies was 20 pM. Equilibrium binding was activated by Ca2+ (KD/Ca2+= 14 nM), inhibited by Mg2+ (IC60= 5.0 mM), and unaffected by AMP-PCP or caffeine. In 200 mM KCI buffer conditions, labeled titration analyses gave only a single site with a KD value similar to and a Bmax value 1.8-fold greater than those obtained for the low-affinity site in 1.0 M KCI buffer. In unlabeled titration measurements, the KD value was fivefold lower, whereas the Bmax value was unaffected. The KD value derived from association and dissociation analysis was 2.4-fold greater in 200 mM KCI compared with 1.0 M KCI buffer conditions. In 200 mM compared with 1.0 M KCI, the potency with which Mg2+ inhibited binding was increased by 3.8-fold, whereas the affinity of the activation site for Ca2+ was reduced by 13-fold. Addition of caffeine in the presence of low salt increased the affinity of Ca2+ activation by 1.7-fold. The inhibitory effect of Mg2+ on [3H]-ryanodine binding in the presence of 200 mM KCI was reversed by AMP-PCP and caffeine with apparent EC50 values of 0.25 and 7.6 mM, respectively. Taken together, these results indicate that ionic strength is an important consideration in binding studies of brain ryanodine receptors and their interactions with modulatory agents.  相似文献   

10.
Circular dichroism has been commonly employed to infer the conformation of DNA in solution. The basis of the conformational assignments is the work of Tunis-Schneider and Maestre, wherein CD spectra of DNA were obtained under conditions comparable to those employed in the x-ray diffraction studies of A-, B-, and C-DNA. It has recently been suggested that the CD spectrum of DNA in chromatin, which is similar to the CD spectrum of the C-form DNA, is a superposition of the normal B-DNA spectrum and a single negative band, centered at 275 nm. This negative band is qualitatively identical to the spectrum for condensed Ψ-form DNA. We have employed the hydrodynamic methods of quasielastic light scattering and sedimentation velocity to determine the extent of DNA tertiary structural alteration in 5.5M LiCl as a possible explanation of the C-form CD spectrum. These studies suggest an eightfold contraction of the Stokes hydrodynamic volume for calf thymus DNA in going from 0.4M NH4Ac to 5.5M LiCl, with no change in molecular weight. The estimated maximum presistence length of DNA in 5.5M LiCl is estimated to be 20.0 nm compared to the “minimum” value of 44.7 nm in NaCl solutions. The value 20.0 nm corresponds to a maximum radius of 16.7 nm for a “continuously coiled” cylinder of DNA, which compares with the value 5.0 nm of DNA in the nucleosome unit of chromatin.  相似文献   

11.
Quasielastic and static light-scattering measurements were made on DNA isolated from chicken erythrocyte mononucleosomes as a function of ionic strength between 6 × 10?4 and 1.0M. A transition from single-exponential autocorrelation functions to markedly non-single-exponential decays was observed around 10?2M ionic strength and was accompanied by a large decrease in the excess light-scattering intensity. Autocorrelation functions recorded below 10?2M salt were well fit by the sum of two exponential relaxation which differed by as much as 100-fold in time constants. Apparent diffusion coefficients for the fast and slow processes plateaued around 10?3M with numerical values approximately 10-fold and 1/10, respectively, of the translational diffusion coefficient for mononucleosome DNA at high ionic strength. This behavior is similar to that observed with poly(L -lysine), for which the slow decay has been associated with a transition to an extraordinary phase. The strong and complex salt dependence observed here illustrates potential difficulties in deriving structural information from scattering by polyions at low ionic strength.  相似文献   

12.
Diffusion of bovine serum albumin in a neutral polymer solution   总被引:3,自引:0,他引:3  
G D Phillies 《Biopolymers》1985,24(2):379-386
The diffusion coefficient D of bovine serum albumin through various solutions (pH 7.0, 0.5M NaCl) of polythylene oxide (Mw ~ 1 × 105, 3 × 105) was studied with quasielastic light scattering. In solutions of the 1 × 105 polymer solution at polymer concentrations above 0.5 g/L, D is considerably greater than would have been expected from the viscosity of water:polymer mixtures, the deviations being larger at low protein concentration that at high protein concentration. With either polymer, D falls with increasing protein concentration.  相似文献   

13.
Y Tsunashima  K Moro  B Chu  T Y Liu 《Biopolymers》1978,17(2):251-265
Group-specific polysaccharides isolated by means of a cetavlon procedure are immunogenic in man and induce protective immunity against meningococcal meningitis. Minute quantities of the polymers in solution can act as vaccines. We now report the first characterization of a fractionated (C-1) group C polysaccharide in 0.4KM KCl and 0.05M sodium acetate by means of light-scattering spectroscopy. Independent measurements of refractive index increments, absolute scattered intensities, angular scattering intensities and line widths as a function of scattering angles and delay times at different concentrations using incident wavelengths of 632.8 nm from a He–Ne laser and of 488 nm from an argon–ion laser yield information on aggregation properties, molecular weight (Mr), radius of gyration 〈r0g1/2z, translational diffusion coefficient 〈D〉0z, and second virial coefficients A2 and B2 of C-1 polysaccharide. At relatively high ionic strength (0.04M KCl + 0.05M sodium acetate), we obtain for the C-1 polysaccharide in solution Mr = 5.15 × 105, 〈r2g1/2z = 345 Å, A2 = 1.25 × 10?4 ml/g, 〈D〉 = 1.16 × 10?7 cm2/sec with a corresponding Stokes radius of 240 Å and B2 = 4.4 ml/g. A2 and B2 are the second virial coefficients from intensity- and diffusion-coefficient measurements. The C-1 polysaccharide aggregates in solution and behaves hydrodynamically like random coils. Viscosity and sedimentation studies further confirm our conclusions that the fractioned C-1 polysaccharide aggregates in solution and EDTA can partially break up those aggregates. However, the system remains polydisperse even after adding an excess amount of EDTA. The weight-average molecular weight of the C-1 polysaccharide in solution depends upon ionic strength and exhibits a minimum at ~0.2M KCl. Finally, viscosity, light-scattering, and sedimentation results all show that the aggregated macromolecular system behaves like random-coiled polymers with no measurable shape factors.  相似文献   

14.
Isothermal microcalorimetry, differential scanning calorimetry (DSC), and chirooptical data obtained for ι-carrageenan in NaCl, LiCl, and NaI aqueous solutions are presented. The experiments have been performed as a function of concentration both for the polymer and for the simple salt as a cosolute. The experimental findings consistently show the occurrence of a salt-induced disorder-to-order transition. From microcalorimetric experiments the exothermic enthalpy of transition ΔHtr is obtained as the difference between the theoretical, purely electrostatic ΔHel enthalpy change and the actual mixing enthalpy ΔHmix, measured when a ι-carrageenan salt-free solution at constant polymer concentration is mixed with a 1:1 electrolyte solution of variable concentration. In the case of added NaCl, the absolute values of enthalpy changes |ΔHtr| are in good agreement with those obtained for the opposite process, at comparable polymer and salt concentrations, from DSC melting curves. The microcalorimetric results show that the negative maximum value of ΔHtr corresponding to the interaction of Li+ counterion with ι-carrageenan polyion results to be significantly lower than the corresponding values obtained for Na+ counterion. At variance with the microcalorimetric data, chirooptical results show that the salt-induced disorder-to-order transition, occurring in the 0.02–0.2M salt concentration range, appears to be complete at a concentration of about 0.08–0.1M of the simple ion, irrespective of the polymer concentration and of the nature of added electrolyte. © 1998 John Wiley & Sons, Inc. Biopoly 45: 105–117, 1998  相似文献   

15.
Mechanochemical study of wet-spun lithium-DNA fibers   总被引:2,自引:0,他引:2  
A Rupprecht 《Biopolymers》1970,9(7):825-842
Uniform LiDNA fiber specimens of nearly 20 m length have been prepared with a wet-spinning method developed by the author. Samples immersed in the spinning bath (80% ethyl alcohol containing 0.4M LiCl) have been subjected to mechanochemical study involving stretching, relaxation, and contraction measurements. A special technique was developed to transfer the sample from the Teflon-coated cylinder used in spinning to the sample column of the mechanochemical apparatus without stretching or removing the sample from the spinning bath. Force–strain curves of samples consisting of two fiber bundles showed an initial region of low slope followed by a region of high slope and a second region of low slope up to rupture. Some thicker specimens showed an aging effect which abolished the initial low-slope region and was interpreted as indicative of crystallization. Force–strain curves of two-bundle samples showed a strong influence of temperature with a complete loss of tensile strength of the LiDNA fibers in the spinning bath at about 55°C. Furthermore, samples at zero strain exhibited a contractile force when subjected to temperatures above about 40°C; the contractile process was pronounced with samples kept above 48°C. On contraction these samples obtained a zero-force length 20–30% of the original. These data are taken as evidence for a helix-to-coil transition occurring in the DNA, the low melting temperature being caused by the chemical influence of the ambient aqueous alcohol–LiCl bath.  相似文献   

16.
The Brownian motion of a single sample of high-molecular-weight poly(L -lysine) [(Lys)n, n = 955] has been studied by dynamic light scattering over a wide range of NaBr concentrations and at three different polyion concentrations. A substantial decrease in scattered intensity is associated with the transition from the ordinary phase to the low-salt extraordinary phase. At the salt concentration where the transition takes place the relaxations are non-exponential and appear to exhibit at all angles a rapid relaxation (τ ? 10 μsec) that is presumed to be a manifestation of the kinetics of the transition process. The K2 dependence of the slow relaxation rates in the extraordinary phase has been confirmed within the experimental error. The extrapolated infinite-dilution values of the diffusion coefficients in the ordinary phase are observed to decline precipitously below 10?2M salt to astonishingly small values, indicating a dramatic rise in the friction factors of the isolated polyions. An extensive discussion of these findings in relation to the theory employed here and to existing data in the literature is also given.  相似文献   

17.
Highly oriented fibers of Li-, Na-, K-, and CsDNA were prepared with a previously developed wet spinning method. The procedure gave a large number of equivalent fiber bundle samples (reference length, L0, typically = 12–15 cm) for systematic measurements of the fiber length L in ethanol–water solutions, using a simple mechanochemical set up. The decrease in relative length L/L0 with increasing ethanol concentration at room temperature gave evidence for the B-A transition centered at 76% (v/v) ethanol for NaDNA fibers and at 80 and 84% ethanol for K- and CsDNA fibers. A smaller decrease in L/L0 of LiDNA fibers was attributed to the B-C transition centered at 80% ethanol. In a second type of experiment with DNA fibers in ethanol–water solutions, the heat-induced helix–coil transition, or melting, revealed itself in a marked contraction of the DNA fibers. The melting temperature Tm, decreased linearly with increasing ethanol concentration for fibers in the B-DNA ethanol concentration region. In the B-A transition region, Na- and KDNA fibers showed a local maximum in Tm. On further increase of the ethanol concentration, the A-DXA region followed with an even steeper linear decrease in Tm. The dependence on the identity of the counterion is discussed with reference to the model for groove binding of cations in B-DNA developed by Skuratovskii and co-workers and to the results from Raman studies of the interhelical bonds in A-DNA performed by Lindsay and co-workers. An attempt to apply the theory of Chogovadze and Frank-Kamenetskii on DNA melting in the B-A transition region to the curves failed. However, for Na- and KDNA the Tm dependence in and around the A-B transition region could be expressed as a weighted mean value of Tm of A- and B-DNA. On further increase of the ethanol concentration, above 84% ethanol for LiDNA and above about 90% ethanol for Na-, K-, and CsDNA, a drastic change occurred. Tm increased and a few percentages higher ethanol concentrations were found to stabilize the DNA fibers so that they did not melt at all, not even at the upper temperature limit of the experiments (~ 80°C). This is interpreted as being due to the strong aggregation induced by these high ethanol concentrations and to the formation of P-DNA. Many features of the results are compatible with the counterion–water affinity model. In another series of measurements, Tm of DNA fibers in 75% ethanol was measured at various salt concentrations. No salt effect was observed (with the exception of LiDNA at low salt concentrations). This result is supported by calculations within the Poisson–Boltzmann cylindrical cell model. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
The proteoglycan subunit (PGS) from bovine nasal cartilage was examined in water and in 0.15 N LiCl by small-angle x-ray scattering (SAXS). The molecular weight of 2.5 × 106 and the radius of gyration, Rg = 493 Å, in 0.15 N LiCl, obtained by SAXS, are in good agreement with values reported by others for similar preparations. Values of the radius of gyration of the cross section, mass per unit length, and persistence length of the PGS are also reported. The low value of intrinsic viscosity ([η]) found in 0.15 N LiCl, and a comparison of the experimental distance distribution function to that of the theoretical distance distribution function for sphere, suggest that the PGS in salt solution approaches spherical symmetry. The much higher value of [η] in water suggests a prolate ellipsoid of low axial ratio.  相似文献   

19.
The reconstitution of Brome Mosaic Virus (BMV) has been studied using neutron scattering. Experiments were performed on disassembled virus without subsequent separation of components. Phase diagrams of the disassembly and subsequent reassembly of BMV were established as a function of pH and LiCl molarity by analytical centrifugation and quasi-elastic light scattering. Disassembly occurs at a pH above 6.5 and above 0.8 M LiCl. On reassembly, if the pH is lowered first, capsids are formed without subsequent incorporation of RNA. Neutron scattering was used to investigate the formation of virus particles, when the ionic strength was lowered from 1.4 to 0.1 M LiCl at pH 7.8. The reconstitution was followed continuously. As it was driven by a lowering of the ionic strength the kinetics of the process cannot be studied for short times. However the fact that at any given ionic strength no evolution of the scattering was observed with time implies that the reconstitution is complete within a few minutes. The observations in buffers with various amounts of D2O lead to the conclusion that the reassembly is achieved by co-condensation of the RNA and of the capsid proteins.  相似文献   

20.
Dynamic light scattering measurements were performed on dilute aqueous solutions of native ovalbumin (OA) and on those of linear OA aggregates induced by thermal denaturation at low ionic strength and neutral pH. The weight-average molecular weight Mw of four aggregates tested ranged from 1,700,000 to 5,500,000. The translational diffusion coefficient D0 of native OA at infinite dilution was estimated as 8.70 × 10 ?7 cm2/s, which gave 56.0 Å as the diameter of the rigid spherical particle. The intensity autocorrelation function of linear OA polymers was analyzed with the cumulant method to obtain the first cumulant Λe. The dependence of Λe on the scattering vector q at very low polymer concentration was found intermediate between those of a flexible chain and a rigid rod. The translational diffusion coefficient Dtr [≡ (Te/q2)q → 0] was in proportion to M, and the magnitude was in good agreement with a value calculated from the wormlike cylinder model with values of three parameters determined in an earlier study, ML = 1600 Å?1, d = 120 Å, and Q = 230 Å, where ML, d, and Q are the molecular weight per unit length, diameter, and persistence length, respectively. Based on these results, a new model, to be called as the dimer model, was proposed to interpret the formation mechanism of linear OA polymers induced by thermal denaturation. © 1993 John Wiley & Sons, Inc.  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