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 15 毫秒
1.
Capillary electrophoresis has been used to determine whether zwitterions contribute to the ionic strength of a solution, by measuring the mobility of a double-stranded DNA oligomer in cacodylate-buffered solutions containing various concentrations of the ionic salt tetraethylammonium chloride (TEA(+)Cl(-)) or the zwitterion tricine(+/-). The mobility of the DNA decreased as the square root of ionic strength, as expected from the Debye-Hückel-Onsager theory of electrophoresis, when TEA(+)Cl(-) was added to the buffer. However, the mobility was independent of the concentration of added tricine(+/-). Hence, zwitterions do not contribute to the ionic strength of a solution.  相似文献   

2.
The free solution mobility of four 20 bp DNA oligomers, with and without A-tracts, has been measured by capillary electrophoresis in Tris-acetate buffer, to test the hypothesis that site-specific binding of monovalent counterions can occur in the narrow minor groove of A-tract DNAs. Preferential counterion binding has been proposed to cause A-tract bending because of asymmetric charge neutralization and collapse of the helix backbone toward the minor groove. Preferential counterion binding in A-tract DNAs should be manifested by a decrease in the electrophoretic mobility observed in free solution, compared to that of non-A-tract DNAs of the same size. Of the four sequences studied here, the slowest absolute mobility, indicative of the greatest counterion binding, was observed for a 20 bp oligomer containing two runs of A3T3 in phase with the helix repeat. A 20-mer containing phased CACA sequences migrated with the fastest mobility; 20-mers containing phased A5 tracts or phased runs of T3A3 migrated with intermediate mobilities. Very similar mobility differences were observed when 1-20 mM NaCl was added to the buffer. The results suggest that preferential counterion binding occurs in A-tract DNAs, especially those containing the AnTn sequence motif.  相似文献   

3.
Pulse gel electrophoresis was used to measure the reduction of mobilities of λ-DNA-Hind III fragments ranging from 23.130 to 2.027 kilobase pairs in Tris borate buffer solutions mixed with either hexammine cobalt(III), or spermidine3+ trivalent counterions that competed with Tris+ and Na+ for binding onto polyion DNA. The normalized titration curves of mobility were well fit by the two-variable counterion condensation theory. The agreement between measured charge fraction neutralized and counterion condensation prediction was good over a relatively wide range of trivalent cation concentrations at several solution conditions (pH, ionic strength). The effect of ionic strength, trivalent cation concentration, counterion structure, and DNA length on the binding were discussed based on the experimental measurements and the counterion condensation theory. © 1996 John Wiley & Sons, Inc.  相似文献   

4.
We have developed a practical analytical treatment of the non-linear Poisson-Boltzmann (P-B) equation to characterize the strong but non-specific binding of charged ligands to DNA and other highly charged macromolecules. These reactions are notable for their strong salt dependence and anti-cooperativity, features which the theory fully explains. We summarize analytical results for concentration profiles and ion binding in various regimes of surface curvature and ionic strength, and show how counterion size and charge distribution may influence competitive binding. We present several practical applications of the formalism, showing how to estimate the ligand concentration needed to effectively compete with a given buffer salt, and how to calculate the amounts of counterion species bound at various distances from the DNA surface under given bulk solution conditions. We cast our results into the form of a Scatchard binding isotherm, showing how the apparent binding constant K(obs) and S = -dlog K (obs )d log[M (+)] can be predicted from the basic theory. Anti-cooperativity arises naturally without steric repulsion, and binding curves can be fitted with K(obs) and effective charge as the only free parameters. We extend the analytical P-B analysis to an arbitrary number of counterion species, and apply the results to fit and predict three-ion competition data.  相似文献   

5.
The free solution electrophoretic mobility of a charged oligomer in an ionic solvent that approximately takes into account relaxation field effects, screening of the velocity field, and the hydrodynamic interactions resulting from motions of the charges due to an electric field is described. For double‐stranded DNA, the free solution electrophoretic mobility under ionic strengths determined by the buffer and pH conditions relevant to capillary electrophoresis increases with increasing molecular weight up to few hundred base pairs. © 1999 John Wiley & Sons, Inc. Biopoly 49: 209–214, 1999  相似文献   

6.
An extension to polyelectrolyte solutions of Onsager's field-dissociation relation for weak electrolytes can be derived in a simple way. It is found that, except in the limit of zero ionic strength, a strong applied electric field prevents counterion condensation from proceeding to completion. The extent of incompleteness initially varies linearly with the applied field. The field-dissociation relation can easily be incorporated into the theory of ionic effects on the stability of ordered polynucleotide Structures, whereupon a dependence of the stability on field strength emerges. An explicit calculation for a cooperative transition of the DNA melting type is presented, and it is concluded that for sufficiently low ionic strengths, a field of the order of 10 kVcm may be able to induce, melting by lowering the Tm by a few degrees. The threshold effect found experimentally by Pörschke, and particularly the observed linear dependence of the threshold field on the logarithm of the ionic strength, appears here as a simple-consequence of the linear increase of the stabilization free energy with the logarithm of ionic strength.  相似文献   

7.
As indicated by recent measurements of the electrophoretic free solution mobility, DNA appears to have a greater helical charge density in Tris-borate-EDTA (TBE) buffers than in Tris-acetate-EDTA (TAE) buffers. Since electrostatic forces play a major role in DNA packaging processes, we have investigated the condensation of closed circular plasmid DNA using total intensity and dynamic light scattering in Tris-borate, Tris-acetate, and Tris-cacodylate buffers with cobaltic hexa-amine (III) [Co(NH(3))(3+)(6)]. We find that neither the critical concentration of Co(NH(3))(3+)(6) nor the hydrodynamic radii of the resulting condensates vary significantly in the buffer systems studied here despite the prediction that DNA condensation should occur at significantly lower Co(NH(3))(3+)(6) concentrations in Tris-borate buffers. Assuming a persistence length behavior similar to B-DNA in the presence of multivalent cations, a decrease in the attractive counterion correlation pressure decay length in Tris-borate buffers does not account for our observations. It is possible that the binding of multivalent cations to DNA may hinder borate association with the DNA double helix.  相似文献   

8.
Mechanical stability of single DNA molecules   总被引:10,自引:0,他引:10       下载免费PDF全文
Using a modified atomic force microscope (AFM), individual double-stranded (ds) DNA molecules attached to an AFM tip and a gold surface were overstretched, and the mechanical stability of the DNA double helix was investigated. In lambda-phage DNA the previously reported B-S transition at 65 piconewtons (pN) is followed by a second conformational transition, during which the DNA double helix melts into two single strands. Unlike the B-S transition, the melting transition exhibits a pronounced force-loading-rate dependence and a marked hysteresis, characteristic of a nonequilibrium conformational transition. The kinetics of force-induced melting of the double helix, its reannealing kinetics, as well as the influence of ionic strength, temperature, and DNA sequence on the mechanical stability of the double helix were investigated. As expected, the DNA double helix is considerably destabilized under low salt buffer conditions (相似文献   

9.
The electrophoresis mobilities of native calf thymus DNA adsorbed on the charged solid particles were measured by a micro-electrophoretic method as functions of pII, ionic strength, and DNA concentration. The mobility data confirm the adsorption of DNA both on the positively charged alumina and negatively charged resin particles at wide range of pH and ionic strength. The mobility data also indicate significant DNA adsorption by negatively charged glass in the acidic range of pH. The electrophoretic mobilities of DNA adsorbed on different substrate particles under identical conditions do not differ widely, indicating the major role of the adsorbed DNA rather than the covered substrate in controlling the charge behavior of the particle. The mobilities of the adsorbed DNA at salt pH are of a comparable order of magnitude to those for the dissolved DNA in solution. The mobility of the adsorbed heat-denatured and alkali-denatured DNA is lower than that of the native adsorbed DNA under identical conditions of pH and ionic strength.  相似文献   

10.
The free solution mobilities of single- and double-stranded DNA molecules of various molecular weights have been measured by capillary electrophoresis in solutions of constant ionic strength containing a common anion and fifteen different monovalent cations. In solutions with the same ionic composition, the mobilities of different DNA molecules can vary by up to 20%, depending on molecular weight, the number of strands, and the presence or absence of A-tracts, runs of four or more contiguous adenine residues. Importantly, the mobilities observed for the same DNA sample can vary by up to 40% in solutions containing different cations. The mobility differences observed for the same DNA in solutions containing different cations cannot be rationalized by differences in the anhydrous radii or intrinsic conductivities of the various cations, or by the sequence-dependent binding of certain cations to A-tracts. Instead, the observed mobilities are linearly correlated with the average number of water-water hydrogen bonds that are present in solutions containing different cations. The mobilities are also correlated with the viscosity B coefficients of the various cations and with the rotational correlation times frictional coefficients observed for water molecules in solutions containing different cations. Hence, monovalent cations modify the free solution mobility of DNA primarily by perturbing the hydrogen-bonded structure of water, affecting the friction experienced by the migrating DNA molecules during electrophoresis.  相似文献   

11.
We present experiments on the bias-induced release of immobilized, single-stranded (ss) 24-mer oligonucleotides from Au-surfaces into electrolyte solutions of varying ionic strength. Desorption is evidenced by fluorescence measurements of dye-labeled ssDNA. Electrostatic interactions between adsorbed ssDNA and the Au-surface are investigated with respect to 1), a variation of the bias potential applied to the Au-electrode; and 2), the screening effect of the electrolyte solution. For the latter, the concentration of monovalent salt in solution is varied from 3 to 1600 mM. We find that the strength of electric interaction is predominantly determined by the effective charge of the ssDNA itself and that the release of DNA mainly occurs before the electrochemical double layer has been established at the electrolyte/Au interface. In agreement with Manning's condensation theory, the measured desorption efficiency (etarel) stays constant over a wide range of salt concentrations; however, as the Debye length is reduced below a value comparable to the axial charge spacing of the DNA, etarel decreases substantially. We assign this effect to excessive counterion condensation on the DNA in solutions of high ionic strength. In addition, the relative translational diffusion coefficient of ssDNA in solution is evaluated for different salt concentrations.  相似文献   

12.
Optically detected thermal stability and biological activity of phage T7 has been compared as the function of the ionic composition and strength of the buffers. The ionic strength range was studied between 20-140 mmol/1. In Tris buffer containing only monovalent ions the biological activity of the phages decreases abruptly below 50 mmol/1 ionic strength. Structural studies show a logarithmic dependence between the ionic strength and the intraphage DNA stability and no significant change in the thermal stability of the whole phage. Mg2+ and Ca2+ ions at low concentration (1 mmol/1) given into a Tris buffer of 20 mmol/1 original ionic strength highly stabilize the biological activity, which stabilization is also to be seen in the intraphage DNA and also in the whole phage thermal denaturation process.  相似文献   

13.
Min KR  Zimmer MN  Rickard AH 《Biofouling》2010,26(8):931-940
The aim of this study was to explore the physicochemical parameters that influence coaggregation between the freshwater bacteria Sphingomonas natatoria 2.1 and Micrococcus luteus 2.13. Using visual coaggregation assays, the effect of different buffers, solutions of differing ionic strength, pH, temperature, and viscosity on the degree of coaggregation was assessed. Coaggregation occurred maximally in distilled water but was inhibited when coaggregates were suspended in a commonly-used oral bacterial coaggregation buffer, saline solutions, and Tris-Cl buffers. Coaggregation was weakly expressed in standard laboratory buffers. The ionic strength of inorganic salt solutions required to inhibit coaggregation depended upon the inorganic salt being tested. Coaggregation occurred at a pH of 3-10, between 5 and 80°C and was inhibited in solutions with a viscosity of 22.5 centipoises at 20°C. Inhibition of coaggregation with NaCl impaired biofilm development. When developing buffers to test for coaggregation, the natural liquid environment should be considered. Coaggregation between S. natatoria 2.1 and M. luteus 2.13 is only affected by physicochemical conditions beyond those typically found in natural freshwater ecosystems. Such a robust ability to coaggregate may enhance the ability of S. natatoria 2.1 and M. luteus 2.13 to develop a niche in freshwater biofilms.  相似文献   

14.
The free solution mobility of DNA has been measured by capillary electrophoresis in the two buffers most commonly used for DNA gel electrophoresis, Tris-borate-EDTA (TBE) and Tris-acetate-EDTA (TAE). The capillaries were coated with polymers of either of two novel acrylamide monomers, N-acryloylaminoethoxyethanol or N-acryloylaminopropanol, both of which are stable at basic pH and effectively eliminate the electroendosmotic mobility due to the capillary walls. The free solution mobility of DNA in TAE buffer was found to be (3.75 ± 0.04) × 10−4 cm2 V−1 s−1 at 25°C, independent of DNA concentration, sample size, electric field strength, and capillary coating, and in good agreement with other values in the literature. The free solution mobility was independent of DNA molecular weight from ∼ 400 base pairs to 48.5 kilobase pairs, but decreased monotonically with decreasing molecular weight for smaller fragments. Surprisingly, the free solution mobility of DNA in TBE buffer was found to be (4.5 ± 0.1) × 10−4 cm2 V−1 s−1, about 20% larger than observed in TAE buffer, presumably because of the formation of nonspecific borate-deoxyribose complexes. © 1997 John Wiley & Sons, Inc. Biopoly 42: 687–703, 1997  相似文献   

15.
We used isothermal titration calorimetry in the temperature range 21-25 degrees C to investigate the effect of pH on the calorimetric enthalpy (delta H(cal)) for sequence specific DNA-binding of the glucocorticoid receptor DNA-binding domain (GR DBD). Titrations were carried out in solutions containing 100 mM NaCl, 1 mM dithiothreitol, 5% glycerol by volume, and 20 mM Tris, Hepes, Mops, or sodium phosphate buffers at pH 7.5. A strong dependence of delta H(cal) on the buffer ionization enthalpy is observed, demonstrating that the DNA binding of the GR DBD is linked to proton uptake at these conditions. The apparent increase in the pK(a) for an amino acid side chain upon DNA binding is supported by the results of complementary titrations, where delta H(cal) shows a characteristic dependence on the solution pH. delta H(cal) is also a function of the NaCl concentration, with opposite dependencies in Tris and Hepes buffers, respectively, such that a similar delta H(cal) value is approached at 300 mM NaCl. This behavior shows that the DNA-binding induced protonation is inhibited by increased concentrations of NaCl. A comparison with structural data suggests that the protonation involves a histidine (His451) in the GR DBD, because in the complex this residue is located close to a DNA phosphate at an orientation that is consistent with a charged-charged hydrogen bond in the protonated state. NMR spectra show that His451 is not protonated in the unbound protein at pH 7.5. The pH dependence in delta H(cal) can be quantitatively described by a shift of the pK(a) of His451 from approximately 6 in the unbound state to close to 8 when bound to DNA at low salt concentration conditions. A simple model involving a binding competition between a proton and a Na(+) counterion to the GR DBD-DNA complex reproduces the qualitative features of the salt dependence.  相似文献   

16.
A common source of error in pH measurements.   总被引:6,自引:0,他引:6       下载免费PDF全文
Glass-electrode assemblies in which the reference half-cell contains a porous ceramic type of liquid junction are likely to produce misleading pH measurements under normal service conditions. The error arises from substantial liquid-junction potentials, associated with the porous ceramic plug, which vary with the ionic composition of the solution under test. The error is not revealed by conventional two-point calibration procedures, since the majority of standard buffer solutions have a similar total ionic strength, but will nevertheless be present when the unknown solution differs in ionic strength from the standardizing buffers. The size of the error is proportional to the ratio between the salt concentration in the standard buffers and the concentration present in the unknown solution, and varies from one electrode specimen to another. The fault was present in 24 out of 30 electrodes in normal use selected at random from seven laboratories, and the mean error was 0.2pH unit per 10-fold salt-concentration difference between standard and test solutions. It is estimated that errors of this order must be widespread in the recent literature. Older pH determinations are likely to be more reliable, since the original reference electrode design with a free-flowing liquid junction is apparently free from the artefact.  相似文献   

17.
The sound velocities in polyelectrolyte solutions were measured at various concentrations of added salts. When aqueous solutions of tetra (n-butyl)ammonium polyacrylate were titrated with concentrated solutions of LiCl, NaCl, KCl or CsCl, the sound velocity, i.e., the adiabatic compressibility of the solution, did not change linearly with added salt concentration, but showed a breaking point. The degrees of counterion binding on polyacrylate ion estimated from the breaking points were 0.25-0.30, independent of cation species. In polystyrenesulfonate, moreover, no Na+ binding was detected from such sound velocity measurements.  相似文献   

18.
The free solution electrophoretic mobility of DNA differs significantly in different buffers, suggesting that DNA-buffer interactions are present in certain buffer systems. Here, capillary and gel electrophoresis data are combined to show that the Tris ions in Tris-acetate-EDTA (TAE) buffers are associated with the DNA helix to approximately the same extent as sodium ions. The borate ions in Tris-borate-EDTA (TBE) buffers interact with DNA to form highly charged DNA-borate complexes, which are stable both in free solution and in polyacrylamide gels. DNA-borate complexes are not observed in agarose gels, because of the competition of the agarose gel fibers for the borate residues. The resulting agarose-borate complexes increase the negative charge of the agarose gel fibers, leading to an increased electroendosmotic flow of the solvent in agarose-TBE gels. The combined results indicate that the buffers in which DNA is studied cannot automatically be assumed to be innocuous.  相似文献   

19.
The interaction of DNA with various neutral pH, amine-based buffers has been analyzed by free solution capillary electrophoresis, using a mixture of a plasmid-sized DNA molecule and a small DNA oligonucleotide as the reporter system. The two DNAs migrate as separate, nearly Gaussian-shaped peaks in 20-80 mM TAE (TAE, Tris-acetate-EDTA; Tris, tris[hydroxymethyl]aminomethane) buffer. The separation between the peaks gradually increases with increasing TAE buffer concentration because of differences in solvent friction between large and small DNA molecules. The two DNAs form complexes with the borate ions in TBE (Tris-borate-EDTA) buffer, with mobilities that depend on the DNA/borate ratio. In 45 mM TBE buffer, the two DNAs comigrate as a single sharp peak, with a mobility that is faster than either of the constituent DNAs in the same buffer. Hence, the mixed DNA-borate complex is stabilized by the binding of additional borate ions, possibly forming bridges between the different DNAs. The mixed DNA-borate complex is gradually dissociated into its component DNAs by increasing the TBE concentration, possibly because the borate binding sites become saturated at high buffer concentrations. Other neutral pH, amine-based buffers, such as Mops (3-[N-morpholino]propanesulfonic acid), Hepes (N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid]), Bes (N,N-bis[2-hydroxyethyl]-2-aminoethanesulfonic acid), Tes (N-tris[hydroxymethyl]methyl-2-aminoethanesulfonic acid), and tricine (N-tris[hydroxymethyl]methylglycine) also form complexes with DNA, giving distorted peaks in the electropherograms. The combined results indicate that borate buffers and most neutral pH, amine-based buffers interact with DNA.  相似文献   

20.
Otim O 《Biopolymers》2001,58(3):329-334
The influence of urea on the viscosity of hydroxyethyl cellulose (HEC), and the state and separation of double-stranded DNA, was studied by viscometry, fluorometry, and capillary electrophoresis. The results show that double logarithm plots of specific viscosity against the volume fraction of HEC in very dilute polymer solutions are linear, the slopes of which decrease from 0.96 in 0M to 0.29 in 7M urea. The linear regression plots converge at 0.0029 g/mL, the entanglement threshold of HEC. The inclusion of urea in HEC solution thus provides an accurate method of determining its entanglement threshold from such plots. Above the entanglement threshold of HEC, urea has no effect on the specific viscosity of HEC. Results also show that urea has no effect on double-stranded DNA. No change in fluorescence was observed when increasing amounts of urea were added to a fixed concentration of DNA. To examine the influence of urea on the migration of DNA in HEC, the separation of DNA was carried out by polymer-solution capillary electrophoresis in HEC solutions containing 0 or 7M urea using unmodified capillary. Observed mobilities were used in data reduction. It was found that a parallel relationship exists between the observed mobilities and the true mobilities. In buffers containing no urea, the pseudo-free solution mobility appears to be independent of the DNA size. It was also observed to be independent of the electric field below 300 V/cm, but relates exponentially to it in 7M urea. The pseudo-retardation constants obtained by Ferguson-like plots were observed to be positive for smaller DNA molecules below 300 V/cm and increasing linearly with electric field in 0M urea, but nearly constant in 7M urea.  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