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.
Proflavine binding may be used as a probe of the environment and interactions of DNA. In this paper we report the effects of the divalent cations Mg++ and putrescine and the trivalent cation spermidine on the proflavine–Na DNA binding equilibrium. Difference spectroscopy at 430 nm was used to determine apparent proflavine–DNA binding constants K at several concentrations of each cation for temperatures between 15 and 43°C, and at a constant total ionic strength of 0.1M. Mg++, putrescine, and spermidine all have greater effects on K than expected on the basis of ionic strength alone in the order spermidine > Mg++ ? putrescine. van't Hoff analysis of K(T) enabled calculation of ΔH° and ΔS°, which are affected differently by each cation. These differences are discussed qualitatively in terms of such concepts as release of condensed counterions, localized or unlocalized condensation, hydration, and restriction of molecular and internal rotation.  相似文献   

2.
Novel palladium(II) complexes ( 7a–7e ) of substituted quinoline derivatives were synthesized. The complexes were characterized using various techniques such as thermogravimetric analysis (TGA), elemental analysis, conductance measurement, mass, absorption, infra‐red (IR), 1H NMR, 13C NMR and energy‐dispersive X‐ray spectroscopy (EDX). Complexes for herring sperm DNA (HS DNA) binding were explored and absorption titration and the binding constant (Kb) as well as Gibb's free energy were evaluated. Complex 7d exhibited the highest binding constant, therefore the thermodynamic parameters of 7d at different temperatures were evaluated. To support the results of the absorption titration, fluorescence titration, viscosity measurement and molecular docking studies were performed. The fluorescence quenching data as evaluated from Stern–Volmer equation were used to calculate KSV, Kf and the number of binding sites. The results of all these studies were in good agreement with the absorption study. DNA electrophoretic mobility was performed to explore the possible application of metal complexes as artificial metallonucleases. The antibacterial activity of the complexes was accessed against different pathogenic bacteria and cytotoxicity was measured using brine shrimp and S. pombe.  相似文献   

3.
Abstract

In this paper, we have studied the in vitro binding of neotame (NTM), an artificial sweetener, with native calf thymus DNA using different methods including spectrophotometric, spectrofluorometric, competition experiment, circular dichroism (CD), and viscosimetric techniques. From the spectrophotometric studies, the binding constant (Kb) of NTM-DNA was calculated to be 2?×?103 M?1. The quenching of the intrinsic fluorescence of NTM in the presence of DNA at different temperatures was also used to calculate binding constants (Kb) as well as corresponding number of binding sites (n). Moreover, the obtained results indicated that the quenching mechanism involves static quenching. By comparing the competitive fluorimetric studies with Hoechst 33258, as a known groove probe, and methylene blue, as a known intercalation probe, and iodide quenching experiments it was revealed that NTM strongly binds in the grooves of the DNA helix, which was further confirmed by CD and viscosimetric studies. In addition, a molecular docking method was employed to further investigate the binding interactions between NTM and DNA, and confirm the obtained results.  相似文献   

4.
Based on equilibrium binding studies, as well as on kinetic investigations, two types of interactions of Cu2+ ions with native DNA at low ionic strength could be characterized, namely, a nondenaturing and a denaturing complex formation. During a fast nondenaturing complex formation at low relative ligand concentrations and at low temperatures, different binding sites at the DNA bases become occupied by the metal ions. This type of interaction includes chelate formation of Cu2+ ions with atoms N(7) of purine bases and the oxygens of the corresponding phosphate groups, chelation between atoms N(7) and O of C(6) of the guanine bases, as well as the formation of specific intestrand crosslink complexes at adjacent G°C pairs of the sequence dGpC. CD spectra of the resulting nondenatured complex (DNA–Cu2+)nat may be interpreted in terms of a conformational change of DNA from the B-form to a C-like form on ligand binding. A slow cooperative denaturing complex formation occurs at increased copper concentrations and/or at increased temperatures. The uv absorption and CD spectra of the resulting complex, (DNA–Cu2+)denat, indicate DNA denaturation during this type of interaction. Such a conclusion is confirmed by microcalorimetric measurements, which show that the reaction consumes nearly the same amount of heat as acid denaturation of DNA. From these and the kinetic results, the following mechanism for the denaturing action of the ligands is suggested: binding of Cu2+ ions to atoms N(3) of the cytosine bases takes place when the cytosines come to the outside of the double helix as a result of statistical fluctuations. After the completion of the binding process, the bases cannot return to their initial positions, and thus local denaturation at the G·C pairs is brought about. The probability of the necessary fluctuations occurring is increased by chelation of Cu2+ ions between atoms N(7) and O of C(6) of the guanine bases during nondenaturing complex formation, which loosens one of the hydrogen bonds within the G·C pairs, as well as by raising the temperature. The implications of the new binding model, which comprises both the sequence-specific interstand crosslinks and the described mechanism of denaturing complex formation, are discussed and some predictions are made. The model is also used to explain the different renaturation properties of the denatured complexes of Cu2+, Cd2+, and Zn2+ ions with DNA. In temperature-jump experiments with the nondenatured complex (DNA–Cu2+)nat, a specific kinetic effect is observed, namely, the appearance of a lag in the response to the perturbation. The resulting sigmoidal shape of the kinetic curves is considered to be a consequence of the necessity of disrupting a certain number of the crosslinks existing in the nondenatured complex before the local unwinding of the binding regions (a main step of denaturing complex formation) may proceed.  相似文献   

5.
The physical (noncovalent) binding of pyrene and phenanthrene to calf-thymus DNA in aqueous NaCl solutions was measured by a spectral method (analysis of absorption spectra by Benesi-Hildebrand plots) and a coupled-column liquid chromatography method (equilibration of DNA solutions with solid hydrocarbon in a generator column and analysis of dissolved hydrocarbon by liquid chromatography). The measurements yielded values of an affinity constant K′ = nK, where n is the apparent number of binding sites per nucleotide and K is the apparent binding constant. The affinity of native DNA for pyrene decreases monotonically with increasing NaCl concentration, whereas the affinity of heat-denatured DNA exhibits a maximum at 0.10M NaCl.K′ for the binding of phenanthrene to native DNA is an order of magnitude lower than K′ for pyrene. The molar enthalpy for the binding of pyrene to native DNA in 10 mM NaCl is (?34.0 ± 1.0) kJ mol?1. The spectral method data indicate that 50 is an upper limit for the average number of base pairs between intercalation sites for pyrene along the DNA helix and that only a fraction of these sites are bound at the highest binding ratios.  相似文献   

6.
I R Miller  D Bach 《Biopolymers》1968,6(2):169-179
The polarographically determined binding constants of Cu++ and Cd++ to DNA were found to decrease with the degree of binding. The binding causes a reduction in the electrostatic potential ψ on the DNA molecule. The binding constant conforms to the relation K = K0 exp {-zeψ/kT}. The binding constant at zero potential K0 depends on the nature of the competing counterfoils to the phosphate group of the DNA; it is apparently smaller when Na+ rather than the ethylpyridonium residue of polyvinylpyridine serves as the competing counterfoil. The specific interact ion between the ethylpyridonium residues of the polybase and the DNA is very weak, even though the polyelectrolytic interaction induced by the decrease of the electrostatic free energy is spectacular. The intereaction is reversible and equilibrium is maintained between the different interaction products. A reshuffling of the interacting partners can be effected by spontaneous or induced (by centrifugal forces) precipitation of the least soluble interaction products participating in the equilibrium.  相似文献   

7.
Abstract

Fourier Transform Infrared (FT-IR) spectra of solid samples of DNA and RNA obtained from freeze-drying at solid CO2 and liquid nitrogen temperatures, have been recorded and correlation between the conformational transitions and spectral changes is proposed. It is concluded that an equilibrium exists between A, B and Z conformations at low temperatures for the DNA molecule, which is temperature dependent, whereas the RNA molecule exhibits only the A conformation. The results have been compared with the metal-adducts of DNA and RNA, where one of the conformations is predominant.

Marker infrared bands for the B conformer have been found to be the strong band at 825 cm?1 (sugar conformer mode) and a band with medium intensity at 690 cm?1 (guanine breathing mode). The A conformation showed characteristic bands at 810 and 675 cm?1. The B to Z conformational transition was characterized by the strong absorption bands near 820-810 cm?1 and at 665-600 cm?1.  相似文献   

8.
The thermal denaturation method was employed to study the effect of Ca2+ and Mn2+ ions on the DNA helix–coil transition parameters at Na+ concentrations of 10?3–10?1M. At low ion concentrations, thermal stability increases, the melting range passes through a maximum, and the denaturation curves become asymmetric. These changes are quantitatively similar for Mn2+ and Ca2+ ions. With a further increase in the concentration of bivalent ions, the conformational transition temperatures pass through a maximum, and the melting range first tends to saturation and then rapidly decreases to 1–2°C. The Mn2+ concentrations, at which the above effects occur, are an order of magnitude lower than the Ca2+ concentrations. Comparison of experimental results and calculation in terms of the ligand theory permitted estimation of binding constants characterizing association between Mn2+ and Ca2+ ions and bases of native and denatured DNA. We show that, unlike the interaction with phosphates, bivalent ion–DNA base binding is weakly dependent on monovalent ion concentration in the solution.  相似文献   

9.
The binding of neomycin sulfate (NS)/paromomycin sulfate (PS) with DNA was investigated by fluorescence quenching using acridine orange (AO) as a fluorescence probe. Fluorescence lifetime, FT-IR, circular dichroism (CD), relative viscosity, ionic strength, DNA melting temperature, and molecular docking were performed to explore the binding mechanism. The binding constant of NS/PS and DNA was 6.70 × 103/1.44 × 103 L mol?1 at 291 K. The values of ΔHθ, ΔSθ, and ΔGθ suggested that van der Waals force or hydrogen bond might be the main binding force between NS/PS and DNA. The results of Stern–Volmer plots and fluorescence lifetime measurements all revealed that NS/PS quenching the fluorescence of DNA–AO was static in nature. FT-IR indicated that the interaction between DNA and NS/PS did occur. The relative viscosity and melting temperature of DNA were almost unchanged when NS/PS was introduced to the solution. The fluorescence intensity of NS/PS–DNA–AO was decreased with the increase in the ionic strength. For CD spectra of DNA, the intensity of positive band at nearly 275 nm was decreased and that of negative band at nearly 245 nm was increased with the increase in the concentration of NS/PS. The binding constant of NS/PS with double-stranded DNA (dsDNA) was larger than that of NS/PS with single-stranded DNA (ssDNA). From these studies, the binding mode of NS/PS with DNA was evaluated to be groove binding. The results of molecular docking further indicated that NS/PS could enter into the minor groove in the A–T rich region of DNA.  相似文献   

10.
The EcoRV restriction endonuclease cleaves DNA at its recognition sequence at least a million times faster than at any other DNA sequence. The only cofactor it requires for activity is Mg2+: but in binding to DNA in the absence of Mg2+, the EcoRV enzyme shows no specificity for its recognition site. Instead, the reason why EcoRV cuts one DNA sequence faster than any other is that the rate of cleavage is controlled by the binding of Mg2+ to EcoRV-DNA complexes: the complex at the recognition site has a high affinity for Mg2+, while the complexes at other DNA sequences have low affinities for Mg2+. The structures of the EcoRV endonuclease, and of its complexes with either 8pecific or non-specific DNA, have been solved by X-ray crystallography. In the specific complex, the protein interacts with the bases in the recognition sequence and the DNA takes up a highly distorted structure. In the non-specific complex with an unrelated DNA sequence, there are virtually no interactions with the bases and the DNA retains a B-like structure. Since the free energy changes for the formation of specific and non-specific complexes are the same, the energy from the specific interactions balances that required for the distortion of the DNA. The distortion inserts the phosphate at the scissile bond into the active site of the enzyme, where it forms part of the binding site for Mg2+. Without this distortion, the EcoRV–DNA complex would be unable to bind Mg2+ and thus unable to cleave DNA. The specificity of the EcoRV restriction enzyme is therefore governed, not by DNA binding as such, but by its ability to organize the structure of the DNA to which it is bound.  相似文献   

11.
This study examines the relationship between the DNA binding thermodynamics and the enzymatic activity of the Klenow and Klentaq Pol I DNA polymerases from Escherichia coli and Thermus aquaticus. Both polymerases bind DNA with nanomolar affinity at temperatures down to at least 5 °C, but have lower than 1% enzymatic activity at these lower temperatures. For both polymerases it is found that the temperature of onset of significant enzymatic activity corresponds with the temperature where the enthalpy of binding (ΔHbinding) crosses zero (TH) and becomes favorable (negative). This TH/activity upshift temperature is 15 °C for Klenow and 30 °C for Klentaq. The results indicate that a negative free energy of DNA binding alone is not sufficient to proceed to catalysis, but that the enthalpic versus entropic balance of binding may be a modulator of the temperature dependence of enzymatic function. Analysis of the temperature dependence of the catalytic activity of Klentaq polymerase using expanded Eyring theory yields thermodynamic patterns for ΔG, ΔH, and TΔS that are highly analogous to those commonly observed for direct DNA binding. Eyring analysis also finds a significant ΔCp of formation of the activated complex, which in turn indicates that the temperature of maximal activity, after which incorporation rate slows with increasing temperature, will correspond with the temperature where the activation enthalpy (ΔH) switches from positive to negative.  相似文献   

12.
Interactions of sulfadiazine sodium (SD‐Na) with calf thymus DNA (ctDNA) and human serum albumin (HSA) were studied using fluorescence spectroscopy, UV absorption spectroscopy and molecular modeling. The fluorescence experiments showed that the processes were static quenching. The results of UV spectra and molecular modeling of the interaction between SD‐Na and ctDNA indicated that the binding mode might be groove binding. In addition, the interaction of SD‐Na with HSA under simulative physiological conditions was also investigated. The binding constants (K) and the number of binding sites (n) at different temperatures (292, 302, 312 K) were 5.23 × 103 L/mol, 2.18; 4.50 × 103 L/mol, 2.35; and 4.08 × 103 L/mol, 2.47, respectively. Thermodynamic parameters including enthalpy change (ΔH) and entropy change (ΔS) were calculated, the results suggesting that hydrophobic force played a very important role in SD‐Na binding to HSA, which was in good agreement with the molecular modeling study. Moreover, the effect of SD‐Na on the conformation of HSA was analyzed using three‐dimensional fluorescence spectra. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

13.
F G Walz  B Terenna  D Rolince 《Biopolymers》1975,14(4):825-837
Spectrophotometric binding studies were undertaken on the interaction of neutral red with native and heat-denatured, sonicated, calf thymus DNA in a 0.2M ionic strength buffer containing Tris–sodium acetate–potassium chloride at 25°C. The pKA of neutral red was found to be 6.81. At pH 5 the binding of protonated neutral red was complicated even at low concentration ratios of dye to DNA. In the pH range 7.5–8.5 the tight binding process could be studied and it was found that both protonated and free base species of neutral red significantly bind with DNA having association constants (in terms of polynucleotide phosphate) of 5.99 × 103 M?1 and 0.136 × 103 M?1, respectively, for native DNA and 7.48 × 103 M?1 and 0.938 × 103 M?1, respectively, for denatured DNA. The pKA value of the neutral red–DNA complexes were 8.46 for native DNA and 7.72 for denatured DNA. These results are discussed in terms of possible binding mechanisms.  相似文献   

14.
The binding of tiamulin with calf thymus DNA was systematically investigated using multispectroscopy and molecular modelling techniques. For DNA, once tiamulin was added, viscosity (η) and melting temperature (Tm) both exhibited an uptrend. The fluorescence performance of the tiamulin–DNA complex did not change with the ionic strength changes. The binding constant (Ka) of tiamulin for single-stranded DNA (ssDNA, 1.48 × 104 M−1) was obviously higher than that for double-stranded DNA (dsDNA, 9.51 × 103 M−1) at 291 K. The helix structure became looser and the base stack force became stronger for DNA due to the presence of tiamulin as seen from circular dichroic (CD) spectra. The intercalation binding mode of tiamulin with DNA was disclosed. Molecular modelling also revealed tiamulin inserting into the base pairs with the lowest binding free energy of −18.73 kJ mol−1 using van der Waals forces as well as hydrogen bonds.  相似文献   

15.
The binding interaction of lovastatin with calf thymus DNA (ct‐DNA) was studied using UV/Vis absorption spectroscopy, fluorescence emission spectroscopy, circular dichroism (CD), viscosity measurement and molecular docking methods. The experimental results showed that there was an obvious binding interaction of lovastatin with ct‐DNA and the binding constant (Kb) was 5.60 × 103 M–1 at 298 K. In the binding process of lovastatin with ct‐DNA, the enthalpy change (ΔH0) and entropy change (ΔS0) were –24.9 kJ/mol and –12.0 J/mol/K, respectively, indicating that the main binding interaction forces were van der Waal's force and hydrogen bonding. The molecular docking results suggested that lovastatin preferred to bind on the minor groove of different B‐DNA fragments and the conformation change of lovastatin in the lovastatin–DNA complex was obviously observed, implying that the flexibility of lovastatin molecule plays an important role in the formation of the stable lovastatin–ct‐DNA complex. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

16.
Interaction of spermine and DNA   总被引:8,自引:0,他引:8  
The effect of spermine upon the denaturation temperature (Tm) of DNA's of various base compositions has been found to depend upon both the base composition of the DNA and the pH of the solution. Measurement of the hydrogen ion titration curve of spermine as a function of temperature reveals that the net charge of the spermine molecule is undergoing a rapid change with temperature in the range of temperatures at which DNA denatures. Since the value of Tm depends upon base composition, the correlation of the effect of spermine upon Tm with the base composition of the DNA used may be explainable in terms of the changing degree of ionization of spermine. The binding of spermine to native DNA has also been studied by dialysis equilibrium. There is no significant variation either in the number of strongly binding sites or strength of binding with base composition. It is concluded that there is no evidence of correlation between the binding of spermine and the base composition of DNA.  相似文献   

17.
Interactions between efavirenz (EFZ) with calf thymus DNA (CT‐DNA) were investigated in vitro under stimulated physiological conditions using multispectroscopic techniques, cyclic voltammetry viscosity measurement, and gel electrophoresis. Methylene blue and acridine orange dyes were used as spectral probes by fluorescence spectroscopy. Hypochromicity was observed in ultra‐violet (UV) absorption band of EFZ. Considerable fluorescence enhancement of EFZ was observed in the presence of increasing amounts of DNA solution and the binding constants (Kf) and corresponding numbers of binding sites (n) were calculated at different temperatures. Thermodynamic parameters including enthalpy change (ΔH) and entropy change (ΔS) were calculated to be –304.78 kJ mol–1 and –924.52 J mol–1 K–1 according to the van ’t Hoff equation, which indicated that reaction is predominantly enthalpically driven. In addition, UV/vis absorption titration of DNA bases confirmed that EFZ interacted with guanine and cytosine preferentially. Gel electrophoresis of DNA with EFZ demonstrated that EFZ also has the ability to cleave supercoiled plasmid DNA. Circular dichroism study showed stabilization of the right‐handed B form of CT‐DNA. All results suggest that EFZ interacts with CT‐DNA via an intercalative mode of binding. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

18.
Studies on poly(L-lysine50, L-tyrosine50)-DNA interaction   总被引:3,自引:0,他引:3  
R M Santella  H J Li 《Biopolymers》1974,13(9):1909-1926
Interaction between poly(Lys50, Tyr50) and DNA has been studied by absorption, circular dichroism (CD), and fluorescence spectroscopy and thermal denaturation in 0.001M Tris, pH 6.8. The binding of this copolypeptide to DNA results in an absorbance enhancement and fluorescence quenching on tyrosine. There is also an increase in the tyrosine CD at 230 nm. The CD of DNA above 250 nm is slightly shifted to the longer wavelength which is qualitatively similar to, but quantitatively much smaller than, that induced by polylysine binding. At physiological pH the poly(Lys50, Tyr50)–DNA complex is soluble until there is one lysine and one tyrosine per nucleotide in the complex. The same ratio of amino acid residues to nucleotide has also been observed in copolypeptide-bound regions of the complex. The addition of more poly(Lys50, Tyr50) to DNA yields a constant melting temperature, Tm′, for bound base pairs at 90°C which is close to that of polylysine-bound DNA under the same condition. The melting temperature, Tm, of free base pairs at about 60°C on the other hand, is increased by 10°C as more copolypeptide is bound to DNA. As the temperature is raised, both absorption and CD spectra of the complexes with high coverage are changed, suggesting structural alteration, perhaps deprotonation, on bound tyrosine. The results in this report also suggest that intercalation of tyrosine in DNA is unlikely to be the mode of binding.  相似文献   

19.
Every method used to quantify biomolecular interactions has its own strengths and limitations. To quantify protein‐DNA binding affinities, nitrocellulose filter binding assays with 32P‐labeled DNA quantify Kd values from 10?12 to 10?8 M but have several technical limitations. Here, we considered the suitability of biolayer interferometry (BLI), which monitors association and dissociation of a soluble macromolecule to an immobilized species; the ratio koff/kon determines Kd. However, for lactose repressor protein (LacI) and an engineered repressor protein (“LLhF”) binding immobilized DNA, complicated kinetic curves precluded this analysis. Thus, we determined whether the amplitude of the BLI signal at equilibrium related linearly to the fraction of protein bound to DNA. A key question was the effective concentration of immobilized DNA. Equilibrium titration experiments with DNA concentrations below Kd (equilibrium binding regime) must be analyzed differently than those with DNA near or above Kd (stoichiometric binding regime). For ForteBio streptavidin tips, the most frequent effective DNA concentration was ~2 × 10?9 M. Although variation occurred among different lots of sensor tips, binding events with Kd ≥ 10?8 M should reliably be in the equilibrium binding regime. We also observed effects from multi‐valent interactions: Tetrameric LacI bound two immobilized DNAs whereas dimeric LLhF did not. We next used BLI to quantify the amount of inducer sugars required to allosterically diminish protein‐DNA binding and to assess the affinity of fructose‐1‐kinase for the DNA‐LLhF complex. Overall, when experimental design corresponded with appropriate data interpretation, BLI was convenient and reliable for monitoring equilibrium titrations and thereby quantifying a variety of binding interactions.  相似文献   

20.
Linker histones (H1) are the basic proteins in higher eukaryotes that are responsible for the final condensation of chromatin. In contrast to the nucleosome core histone proteins, the role of H1 in compacting DNA is not clearly understood. In this study ITC was used to measure the binding constant, enthalpy change, and binding site size for the interactions of H10, or its C-terminal (H10-C) and globular (H10-G) domains to highly polymerized calf-thymus DNA at temperatures from 288 K to 308 K. Heat capacity changes, ΔCp, for these same H10 binding interactions were estimated from the temperature dependence of the enthalpy changes. The enthalpy changes for binding H10, H10-C, or H10-G to CT-DNA are all endothermic at 298 K, becoming more exothermic as the temperature is increased. The ΔH for binding H10-G to CT-DNA is exothermic at temperatures above approximately 300 K. Osmotic stress experiments indicate that the binding of H10 is accompanied by the release of approximately 35 water molecules.We estimate from our naked DNA titration results that the binding of the H10 to the nucleosome places the H10 protein in close contact with approximately 41 DNA bp. The breakdown is that the H10 carboxyl terminus interacts with 28 bp of linker DNA on one side of the nucleosome, the H10 globular domain binds directly to 7 bp of core DNA, and shields another 6 linker DNA bases, 3 bp on either side of the nucleosome where the linker DNA exits the nucleosome core.  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