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.
Monomolecular condensation of lambda-DNA induced by cobalt hexamine   总被引:9,自引:0,他引:9  
J Widom  R L Baldwin 《Biopolymers》1983,22(6):1595-1620
Measurements of static and dynamic light scattering have been used to distinguish between monomolecular DNA condensation and aggregation of condensed molecules. In low salt, using Co3+(NH3)6 as the condensing agent, and at λ-DNA concentrations below 0.2 μg/mL, the transition curves for monomolecular condensation and aggregation are well separated for times of 16 h. In these conditions, the intensity of scattered light (90°) and also the diffusion coefficient of the condensed DNA show reasonable values for monomolecular condensation that are independent of DNA concentration and also of Na+ Co3+(NH3)6 concentrations for which monomolecular condensation is complete. At higher Co3+(NH3)6 concentrations, which produce aggregation (as judged by the intensity of scattered light), the diffusion coefficient decreases sharply. The transition curve for monomolecular condensation is independent of DNA concentration but shows a hysteresis loop. The kinetics of condensation are slow in the forward direction and fast in the reverse direction, indicating that the actual transition curve is measured closely by reversal experiments. Aggregation is blocked kinetically in both the forward and reverse directions when Co3+(NH3)6 is the condensing agent at low Na+ concentrations. When spermine or spermidine is the condensing agent and observations are made at 16 h, it is not possible to separate the transition curves for monomolecular condensation and for aggregation in conditions that are successful with Co3+(NH3)6. Some interesting properties of monomolecular condensation are noted. (1) The transition is not a two-state reaction, as judged by measurements of the diffusion coefficient through the transition zone. (2) The transition for monomolecular condensation is diffuse. (3) The dimensions of the monomolecular condensates have been calculated from the translational diffusion coefficient for an assumed toroidal shape by the formula derived by Allison and coworkers [(1981) Biopolymers 20 , 469–488]. These dimensions are in reasonable agreement with ones deduced from electron microscopy by Chattoraj and coworkers [(1978) J. Mol. Biol. 121 , 327–337]. (4) The phase diagram relating the Na+ to the Co3+(NH3)6 concentrations needed for condensation has a slope of 0.6 in a log–log plot. According to numerical solutions of Manning's theory for the atmospheric binding of competing cations to DNA, this means that condensation occurs at a late stage in the replacement of Na+ by Co3+(NH3)6 around the DNA. The fraction of DNA phosphate charge neutralized at condensation is computed to be in the neighborhood of 0.9, as found by Wilson and Bloomfield [(1979) Biochemistry 18 , 2192–2196], but to vary with the Na+ concentration.  相似文献   

2.
Static and dynamic light scattering measurements were made of solutions of pGem1a plasmids (3730 base pairs) in the relaxed circular (nicked) and supercoiled forms. The static structure factor and the spectrum of decay modes in the autocorrelation function were examined in order to determine the salient differences between the behaviors of nicked DNA and supercoiled DNA. The concentrations studied are within the dilute regime, which is to say that the structure and dynamics of an isolated DNA molecule were probed. Static light scattering measurements yielded estimates for the molecular weight M, second virial coefficient A2, and radius of gyration RG. For the nicked DNA, M = (2.8 ± 0.4) × 106g/mol, A2 = (0.9 ± 0.2) × 10−3 mol cm3/g2, and RG = 90 ± 3 nm were obtained. For the supercoiled DNA, M = (2.5 ± 0.4) × 106 g/mol, A2 = (1.2 ± 0.2) × 10−3 mol cm3/g2, and RG = 82 ± 2.5 nm were obtained. The static structure factors for the nicked and supercoiled DNA were found to superpose when they were scaled by the radius of gyration. The intrinsic stiffness of DNA was evident in the static light scattering data. Homodyne intensity autocorrelation functions were collected for both DNAs at several angles, or scattering vectors. At the smallest scattering vectors the probe size was comparable to the longest intramolecular distance, while at the largest scattering vectors the probe size was smaller than the persistence length of the DNA. Values of the self-diffusion coefficients D were obtained from the low-angle data. For the nicked DNA, D = (2.9 ± 0.3) × 10−8 cm2/s, and for the supercoiled DNA, D = (4.11 ± 0.21) × 10−8 cm2/s. The contribution to the correlation function from the internal dynamics of the DNA was seen to result in a strictly bimodal decay function. The rates of the faster mode Γint, reached plateau values at low angles. For the nicked DNA, Γint = 2500 ± 500 s−1, and for the supercoiled DNA, Γint = 5000 ± 500 s−1. These rates correspond to the slowest internal relaxation modes of the DNAs. The dependence of the relaxation rates on scattering vector was monitored with the aid of cumulants analysis and compared with theoretical predictions for the semiflexible ring molecule. The internal mode rates and the dependence of the cumulants moments reflected the difference between the nicked DNA and the supercoiled DNA dynamical behavior. The supercoiled DNA behavior seen here indicates that conformational dynamics might play a larger role in DNA behavior than is suggested by the notion of a branched interwound structure. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
A very low-angle light-scattering photometer is described with respect to optical features, scattering cell, correction factors, and absolute calibration in the angular range 2°–35°. An improved microfiltration apparatus was used to obtain essentially dust-free aqueous solutions for very low-angle light scattering. The instrument was calibrated with silicotungstic acid, an absolute molecular-weight standard, and the calibration was confirmed with the use of several secondary standards. Very low-angle light-scattering measurements were made to determine the weight-average molecular weight M?r and z-average radius of gyration Rg,z of a commerical preparation of calf-thymus DNA. Microfiltration of the solutions allowed measurements down to 6°. The value M?r = 20.0 × 106 obtained by extrapolating 6°–9° data to 0° is more than three times that from 30°–75° data (6.38 × 106) but ~20% smaller than that from 10–35° data (23.7 × 106). The experimental errors in M?r and Rg,z are estimated to be ±8% and ±14%, respectively. Combined 6°–75° data from two photometers fit well a theoretical scattering curve for a model wormlike coil of the same M?r as the DNA sample.  相似文献   

4.
The instability of aqueous solutions of poly(α-L -glutamic acid) (PGA) at low pH is due to two distinguishable phenomena: precipitation, favored above 40°C., and aggregation, favored below 20°C. The aggregated form of PGA can be isolated by gel permeation chromatography. Both aggregation and precipitation increase with decreasing pH, i.e., with decreasing ionization of the side chain carboxyl groups. Temperature-induced aggregation and disaggregation give rise to a reproducible hysteresis loop which can be followed by optical rotation, light scattering, sedimentation, viscosity, and chromatography. Hysteresis has been observed with different PGA samples, and in several aqueous buffered or unbuffered solvents and organic-aqueous solvent mixtures and in the pH range 4.1–4.5. Aggregation manifests itself as an increase in negative optical rotation in the visible and ultraviolet spectral range. The specific relation at 233 mμ is sensitive to aggregation and also reflects the hysteresis. Measurements of optical rotatory dispersion indicate that a0 reflects the hysteresis but b0 does not, the latter revealing only reversible changes with aggregation and disaggregation. The helix-coil equilibrium is apparently unperturbed by aggregation, as is the thermal stability of the helix structure. For fully aggregated PGA it is estimated that a0 increases by about 300 degrees, which suggests that a0 may be a sensitive parameter to measure aggregation in other systems. The rate of aggregation increases with decreasing temperature. The disaggregation, upon heating, is more rapid. However, kinetics measurements have not yet been done. The temperature M at which all aggregates are disrupted increases with decreasing pH, but is independent of total PGA concentration, at constant pH. No molecular weight dependence of M was detected in the range 20–100 × 103. The shape and size of the hysteresis loop depends upon pH and molecular weight, which is interpreted as a dependence on the extent of aggregation. One branch of the loop, representing the helix–coil transition of isolated molecules, is reversible, while the others, representing the formation and disruption of the aggregates, are not. The system exhibits both ascending and descending scanning curves, which are typical of a true hysteresis.  相似文献   

5.
The secondary structure of xanthan in solutions of relatively low salt concentration and at room temperature has been investigated using static light scattering experiments. Additional evidence has been found for a dimeric structure at 25°C in 0.01M NaCl. From the experimental z-average mean square (ms) radius of gyration, a value for the persistence length p has been estimated, taking explicitly into account the polydispersity of the three samples used, which has been established by gel permeation chromatography (GPC) measurements. The experimental particle scattering functions of the three samples are consistent with theoretical estimates for polydisperse systems with the same value of p = 65 ± 10 nm and the molar mass per unit length for a dimeric structure. This secondary structure remains unaffected by the ionic strength in the 0.005–0.0lM range. Partial aggregation seems to occur at higher NaCl concentrations. Light scattering and GPC data show that heating the xanthan 0.01M NaCl solutions to about 70°C considerably reduces the Mw of the low molar mass sample (2.3 × 105-g·mol?1), contrary to what is observed for the high molar mass sample (1.8 × 106-g·mol?1). These experimental findings can be accounted for by a partial temperature-induced dissociation of the xanthan dimers according to an all-or-none mechanism. © 1994 John Wiley & Sons, Inc.  相似文献   

6.
Purified recA protein is induced by high salt concentrations to hydrolyse ATP even in the absence of DNA. By small angle neutron scattering we show that this salt activation results from a structural transition of the protein filament in the presence of ATPγS from the inactive, compact form (a helical polymer of pitch 70 Å and cross-sectional radius of gyration Rc 40 Å) to the open form (a helical filament of pitch 95 Å and Rc 35 Å, which are the same structural parameters as in the ATPase active complex with DNA and ATP), without detectable change in the degree of association. We conclude that activation of recA is due to the same structural change whether induced by the binding of DNA or by salt. Indeed, the other enzymatic activity of recA, the proteolytic cleavage of the lexA repressor, is found to be inducible by the same salt concentrations as those of the structural transition.  相似文献   

7.
Thermodynamics of the B to Z transition in poly(dGdC)   总被引:1,自引:0,他引:1  
The thermodynamics of the B to Z transition in poly(dGdC) was examined by differential scanning calorimetry, temperature-dependent absorbance spectroscopy, and CD spectroscopy. In a buffer containing 1 mM Na cacodylate, 1 mM MgCl2, pH 6.3, the B to Z transition is centered at 76.4°C, and is characterized by ΔHcal = 2.02 kcal (mol base pair)?1 and a cooperative unit of 150 base pairs (bp). The tm of this transition is independent of both polynucleotide and Mg2+ concentrations. A second transition, with ΔHcal = 2.90 cal (mol bp)?1, follows the B to Z conversion, the tm of which is dependent upon both the polynucleotide and the Mg2+ concentrations. Turbidity changes are concomitant with the second transition, indicative of DNA aggregation. CD spectra recorded at a temperature above the second transition are similar to those reported for ψ(–)-DNA. Both the B to Z transition and the aggregation reaction are fully and rapidly reversible in calorimetric experiments. The helix to coil transition under these solution conditions is centered at 126°C, and is characterized by ΔHcal = 12.4 kcal (mol bp)?1 and a cooperative unit of 290 bp. In 5 mM MgCl2, a single transition is seen centered at 75.5°C, characterized by ΔHcal = 2.82 kcal (mol bp)?1 and a cooperative unit of 430 bp. This transition is not readily reversible in calorimetric experiments. Changes in turbidity are coincident with the transition, and CD spectra at a temperature just above the transition are characteristic of ψ(–)-DNA. A transition at 124.9°C is seen under these solution conditions, with ΔHcal = 10.0 kcal (mol bp)?1 and which requires a complex three-step reaction mechanism to approximate the experimental excess heat capacity curve. Our results provide a direct measure of the thermodynamics of the B to Z transition, and indicate that Z-DNA is an intermediate in the formation of the ψ-(–) aggregate under these solution conditions.  相似文献   

8.
Inhibition of cation-induced DNA condensation by intercalating dyes   总被引:4,自引:0,他引:4  
J Widom  R L Baldwin 《Biopolymers》1983,22(6):1621-1632
Several intercalating dyes are shown to inhibit the cation-induced condensation of λ-DNA when Co3+(NH3)6 is the condensing agent. The dyes that have been studied are ethidium, propidium, proflavin, quinacrine, and actinomycin D. Earlier work has shown that intercalating dyes inhibit ψ-DNA condensation. [Lerman, L. S. (1971) Prog. Mol. Subcell. Biol. 2 , 382–391; Cheng, S. & Mohr, S. C. (1975) Biopolymers 14 , 663–674.] Dye-induced decondensation of intramolecularly condensed DNA has been studied by making use of conditions in which Co3+(NH3)6 produces intramolecular condensation without significant aggregation. Some aggregation is caused, however, during dye-induced decondensation. Dye titration curves of DNA decondensation have been measured by excess light scattering to monitor decondensation and by fluorescence to monitor intercalation. All of the dyes studied act as competing cations in displacing the condensing cation Co3+(NH3)6 from the DNA. Competition occurs both in and below the transition zone for condensation. The effectiveness of a dye as a competing cation increases with its net positive charge. Before decondensation begins, no intercalated dye can be detected, suggesting that intercalation might be incompatible with the proper helix packing needed for cation-induced DNA condensation. To test this last point, methidium–spermine was synthesized: it contains an intercalating methidium head group combined with a polyamine tail. Methidium–spermine is found to cause λ-DNA condensation, but aggregation accompanies condensation, as has been found earlier for spermine and spermidine. Fluorescence and absorption spectra indicate that the methidium group is intercalated when the DNA is condensed, indicating that intercalation need not be incompatible with DNA condensation. The presence of aggregates among the condensed DNA molecules makes this last conclusion tentative.  相似文献   

9.
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   

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

11.
At low salt ([Na+] = 10?3M), spermidine is capable of transforming DNA from a highly extended random coil to a compact particle. The transition takes place at a spermidine concentration of around 25 μM and the compact particle has been previously studied in considerable detail for several different DNAs. The objective of the present study is to see what effect, if any, spermidine has on T7 DNA conformation prior to collapse using flow dichroism and intrinsic viscosity. We conclude that increasing the spermidine concentration from 0 to the collapse transition point (above 20 μM) makes DNA increasingly nondraining. Furthermore, the persistence length dropped from 785 (±42) to 560 (±32) to 445 (±26) Å on increasing the ambient spermidine concentration from 0 to 1 to 10 μM. These results are in good agreement with counterion condensation theory and Odijk's theory of the electrostatic contribution to the persistence length of DNA. Nonetheless, it is concluded that counterion condensation is not entirely responsible for DNA collapse and that crosslinking promotes the transition to the compact state.  相似文献   

12.
The relatively novel combination of flow field-flow fractionation (FFF) and multiangle light scattering (MALS) was employed to study a nondegraded κ-carrageenan in different 0.1M salt solutions. The applicability of the technique was tested, and the effects of salt type and salt composition on the molar mass and radius of gyration were studied. A conformational ordering was induced at room temperature by switching the solvent from 0.1M NaCl (coil form) to 0.1M NaI (helix form). An approximate doubling of the average molar mass and an increase in radius of gyration was then observed, in agreement with results obtained previously using size exclusion chromatography–MALS. This increase in size was attributed to conformational ordering and to the formation of double helices. Severe aggregation was observed above 40% CsI in the 0.1M mixed salt solution of CsI and NaI. This was ascribed to the association of helices into large aggregates. For these large associates, having molar masses of several millions, a reversal of the elution order in flow FFF was detected. © 1998 John Wiley & Sons, Inc. Biopoly 45: 85–96 1998  相似文献   

13.
The DNA helix–coil transition has been studied in the presence of high concentrations of manganese ions (about 10?3M), which corresponds to the conditions close to equal stability of the A+T and G+C pairs, at the ionic strengths of 10?1, 10?2, and 1.6 × 10?3M Na+. With the Mn2+ ion effect, the transition range is significantly reduced to not more than 0.2°C at 1.2 × 10?3M Mn2+ and 1.6 × 10?3M Na+. The melting curves display a sharp kink at the end of the helix–coil transition, which is interpreted as an indication of the second-order phase transition. It is shown that the melting curves obtained can be approximated by a simple analytical expression 1 – θ = exp[–a(tc - t)], where θ is the DNA helix fraction, tc is the phase transition temperature, and a is an empirical parameter characterizing the breadth of the melting range and responsible for the magnitude of a jump of the helicity derivative with respect to the temperature at the phase transition point.  相似文献   

14.
15.
Despite the extensive interest in applications of plasmid DNA, there have been few direct measurements of the root mean square radius of gyration, RG, of different plasmid isoforms over a broad range of plasmid size. Static light scattering data were obtained using supercoiled, open‐circular, and linear isoforms of 5.76, 9.80, and 16.8 kbp plasmids. The results from this study extend the range of RG values available in the literature to plasmid sizes typically used for gene therapy and DNA vaccines. The experimental data were compared with available theoretical expressions based on the worm‐like chain model, with the best‐fit value of the apparent persistence length for both the linear and open‐circular isoforms being statistically identical at 46 nm. A new expression was developed for the radius of gyration of the supercoiled plasmid based on a model for linear DNA using an effective contour length that is equal to a fraction of the total contour length. These results should facilitate the development of micro/nano‐fluidic devices for DNA manipulation and size‐based separation processes for plasmid DNA purification. Biotechnol. Bioeng. 2010;107: 134–142. © 2010 Wiley Periodicals, Inc.  相似文献   

16.
G Bressan  R Rampone  E Bianchi  A Ciferri 《Biopolymers》1974,13(11):2227-2240
Optical density, viscosity, and light scattering measurements for calf thymus DNA in water–N,N dimethyl formamide (DMF) solutions are presented. DMF content varied between 0 and 60% (v/v) and DNA molecular weight varied between 15 × 106 and 0.5 × 106. Complementary measurements of the solubility of adenine, thymine, guanine, and cytosine in H2O–DMF mixtures are presented. The denaturation temperature of DNA, manifested by about a 35% increase of optical density, is gradually depressed by increasing DMF content. However, a significant increase of OD occurs even before (and even after) the denaturation point, when DMF content is increased isothermally. The intrinsic viscosity also exhibits a large decrease when DMF content is increased both before and after the denaturation point. Light scattering data for high-molecular-weight DNA in the predenaturation range indicate a decrease of the mean-square radius and a constant molecular weight on increasing DMF content. The results, interpreted in terms of the wormlike chain of Kratky and Porod, indicate a large decrease of the persistence length of DNA. For low-molecular-weight DNA, radius and molecular weight increase with DMF content, indicating intermolecular aggregation. The formation of compact structures of native DNA is discussed in terms of an increased solubility of uncharged bases, and a decreased solubility of phosphate and deoxyribose groups, when a less polar environment is provided by the addition of DMF.  相似文献   

17.
Malate synthase was investigated in solution by the small-angle X-ray scattering technique. The substrate-free enzyme was shown to have a molecular weight of 186000, a radius of gyration of 3.96 nm, a maximum particle diameter of 11.2 nm, a volume of 343 nm3, a radius of gyration of the thickness of 1.04 nm, and an axial ratio of 1:0.33. The enzyme molecule undergoes small changes in overall structure upon binding substrates. Investigation of the enzyme under prolonged exposure to X-rays led to an aggregation of the enzyme and allowed statements concerning the way of aggregation and factors influencing aggregation.  相似文献   

18.
19.
20.
Morphological and conformational transitions of lentinan (LT), a β-glucan extracted from Shiitake mushrooms (Lentinula edodes), were investigated at different concentrations of aqueous NaOH, using Small Angle X-ray Scattering (SAXS) technique. At low NaOH(aq) concentration LT chains are self-associated and adopt the triple helix form where as at higher NaOH concentrations the polymer chains undergo a transition to random coil chains. Also, the presence of fractal dimensions was observed through analysis of the exponential decay of the scattering intensity as a function of the scattering angle. In addition, the lateral radius of gyration was determined for LT in different concentrations of NaOH solution, indicating a rigid triple helix present as a small rod-like structure. Interactions of LT with two zwitterionic surfactants were investigated by surface tension, fluorescence, and static light scattering measurements. Experimental data showed that the formation of LT–(surfactant) complexes occurred through a cooperative process.  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