首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Single B-form DNA molecules undergo an overstretching transition at force Fov to a ~1.7-fold longer form when stretched. The nature of overstretched DNA has been debated for over 10?years. Either peeled (PL DNA), internally melted (M DNA), or unwound double-helical (S DNA) forms of overstretched DNA have been suggested. Here, we characterize the kinetics of the overstretching transition in polymeric torsionally unconstrained double-stranded (ds) DNA molecules. We pull ~50?Kbp λ–DNA molecules using optical tweezers with rates ν ~10?nm/s to 5?×?104?nm/s, (overstretching time between 0.2 and 103?s). The Fov(ν, [Na+]) dependence measured over a broad range of rates and solution ionic strength suggests the existence of all three forms of the overstretched DNA. Thus, at [Na+]?>?50?mM and the stretching time >>1?s, internal melting dominates overstretching. This B-to-M transition is highly cooperative (involves ~100?bp), and slow (on/off time ~1000?s). Faster overstretching during ?1?s leads to B-to-S DNA transition, which is less cooperative (involves ~10?bp) and faster (on/off time ~1?s). In contrast, in lower salt ([Na+]?<?50?mM), the overstretching during >1?s leads to DNA peeling. However, on the faster time scale of 0.2–1?s, even in low salt, the DNA overstretches into S DNA, as peeling becomes kinetically prohibited. Our conclusions are supported by several independent lines of evidence, including the salt and rate dependence of both the slope of the overstretched DNA force-extension curve and the value of the second transition force (from M or PL DNA into S DNA).  相似文献   

2.
Abstract

The interaction between Tb(IV)-NR complex and herring sperm DNA in buffer solution of Tris-HCl was investigated with the use of acridine orange(AO) as a spectral probe. The binding modes and other information were provided by the UV–spectrophotometry and fluorescence spectroscopy. The thermodynamic functions expressed that the binding constants of Tb(IV)-NR complex with DNA was Kθ298.15K = 4.03?×?105?L·mol?1, Kθ310.15K =1.30?×?107?L·mol?1, and the ΔrGθ m 298.15?K=?3.20?×?104 J·mol?1. The scatchard equation suggested that the interaction mode between Tb(IV)-NR complex and herring sperm DNA is electrostatic and weak intercalation bindings. FTIR spectroscopy results also indicate that there is a specific interaction between the Tb(IV)-NR complex and the A and G bases of DNA.  相似文献   

3.
In this study, an attempt has been made to study the interaction of a Zn(II) complex containing an antibiotic drug, ciprofloxacin, with calf thymus DNA using spectroscopic methods. It was found that Zn(II) complex could bind with DNA via intercalation mode as evidenced by: hyperchromism in UV–Vis spectrum; these spectral characteristics suggest that the Zn(II) complex interacts with DNA most likely through a mode that involves a stacking interaction between the aromatic chromophore and the base pairs of DNA. DNA binding constant (Kb = 1.4 × 104 M?1) from spectrophotometric studies of the interaction of Zn(II) complex with DNA is comparable to those of some DNA intercalative polypyridyl Ru(II) complexes 1.0 ?4.8 × 104 M?1. CD study showed stabilization of the right-handed B form of DNA in the presence of Zn(II) complex as observed for the classical intercalator methylene blue. Thermodynamic parameters (ΔH < 0 and ΔS < 0) indicated that hydrogen bond and Van der Waals play main roles in this binding prose. Competitive fluorimetric studies with methylene blue (MB) dye have shown that Zn(II) complex exhibits the ability of this complex to displace with DNA-MB, indicating that it binds to DNA in strong competition with MB for the intercalation.  相似文献   

4.
The interaction of the Trp–Sm(III) complex with herring sperm DNA (hs‐DNA) was investigated with the use of acridine orange (AO) dye as a spectral probe for UV‐vis spectrophotometry and fluorescence spectroscopy. The results showed that the both the Trp–Sm(III) complex and the AO molecule could intercalate into the double helix of the DNA. The Sm(III)–(Trp)3 complex was stabilized by intercalation into the DNA with binding constants: K?25°C = 7.14 × 105 L·mol?1 and K?37°C = 5.28 × 104 L·mol?1, and it could displace the AO dye from the AO–DNA complex in a competitive reaction. Computation of the thermodynamic functions demonstrates that ΔrHm? is the primary driving power of the interaction between the Sm(III)(Trp)3 complex and the DNA. The results from Scatchard and viscometry methods suggested that the interaction mode between the Sm(III)(Trp)3 complex and the hs‐DNA is groove binding and weak intercalation binding. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

5.
The toxic interaction of melamine with herring sperm DNA (hs‐DNA) was investigated by using fluorescence and UV–vis absorption spectra techniques. The experimental results showed that the toxic interaction between melamine and hs‐DNA occurred. Fluorescence quenching experiments indicated the existence of electrostatic binding between melamine and hs‐DNA. The binding constants KA and the binding site numbers were calculated by means of the Stern–Volmer equation and were 9.8 × 104 L mol?1 and 1.3, respectively. Both the results of fluorescence spectra and UV–vis absorption spectra verified that there are electrostatic binding between melamine and hs‐DNA. The possibility in the presence of a classical intercalation binding mode could be ruled out by using DNA unwinding experiments. © 2010 Wiley Periodicals, Inc. J Biochem Mol Toxicol 24:323–329, 2010; View this article online at wileyonlinelibrary.com . DOI 10.1002/jbt.20341  相似文献   

6.
The influence of water-soluble cationic meso-tetra-(4?N-oxyethylpyridyl)porphyrin (H2TOEPyP4) and it’s metallocomplexes with Ni, Cu, Co, and Zn on hydrodynamic and spectral behavior of DNA solutions has been studied by UV/Vis absorption and viscosity measurement. It was shown that the presence of planar porphyrins such as H2TOEPyP4, NiTOEPyP4, and СuTOEPyP4 leads to an increase in viscosity at relatively small concentrations, and then decrease to stable values. Such behavior is explained by intercalation of these porphyrins in DNA structure because the intercalation mode involves the insertion of a planar molecule between DNA base pairs which results in a decrease in the DNA helical twist and lengthening of the DNA. Further decrease of viscosity is explained by the saturation intercalation sites and occurs outside the binding mode. But, in the case of porphyrins with axial ligands such as CoTOEPyP4 and ZnTOEPyP4, the hydrodynamic parameters decrease, which is explained by self-stacking of these porphyrins in DNA surface. This data are proved by spectral measurements. The results obtained from titration experiments were used for calculation of binding parameters: the binding constant K b and the number of binding sites per base pair n. Obtained data reveal that K b varies between 3.4 and 5.4?×?106?M?1 for a planar porphyrins, a range typical for intercalation mode interactions, and 5.6?×?105?M?1 and 1.8?×?106?M?1 for axial porphyrins. In addition, the exclusion parameter n also testifies that at intercalation, (n~2) the adjacent base pairs are removed to place the planar molecules, and for outside binders to pack on the surface needs too few places (n~0.5–1). It is apparent that the binding is somewhat stronger at intercalation. The viscometric and spectrophotometric measurements are in good agreement.  相似文献   

7.
Using 1-6-12 atom-atom potential functions with a solvent-averaged electrostatic potential U = qIqj/(?D(r)r1j between the charges qI and qj the stereospecific interaction of Na+ with clusters of l-glycero-phosphorylethanolamine (GPE) and -choline (GPC) in a two-dimensional planar lattice was calculated. With neutral PE headgroups only a weak attractive interaction of ?140 kJ/mol was found. This should be compensated in PE/water systems by reorientational processes within the Na+ hydration shell to get an optimum binding of Na+ with the PE headgroup. With a phosphorylethanolamine anion within a neutral PE headgroup environment a stereospecific binding energy of ?500 kJ/mol was calculated. The interaction of Na+ or Ca2+ with clusters of phosphatidylcholine is always repulsive. These results are in a very good agreement with experimental ones where a binding of Na+ with PE bilayers at pH > 7.5, but not with PC bilayers was detected.  相似文献   

8.
The interaction between DNA and furazolidone/furacillin was investigated using various analytical techniques including spectroscopy and electroanalysis and molecular modelling. With the aid of acridine orange (AO), the fluorescence lifetimes of DNA–AO, DNA–furazolidone/furacillin–AO remained almost the same, which proved that the ground state complex was formed due to furazolidone/furacillin binding with DNA. Circular dichroism spectra and Fourier transform infrared spectroscopy showed that the second structure of DNA changed. Viscosity experiments presented that relative viscosity of DNA was increased with the increasing concentrations of furazolidone and almost unchanged for furacilin. In addition, the results of melting temperature (Tm), ionic strength, site competition experiments, cyclic voltammetry, and molecular docking all proved the intercalation binding mode for furazolidone and groove binding mode for furacilin. The binding constants (Ka) obtained from Wolfe–Shimmer equation were calculated as 3.66 × 104 L mol?1 and 3.95 × 104 L mol?1 for furazolidone–DNA and furacilin–DNA, respectively.  相似文献   

9.
Abstract

The interaction of the [Mn(mef)2(phen)H2O] complex in which mef is mefenamic acid drug and phen is 1,10 phenanthrolin ligand with calf thymus DNA (ct-DNA) was studied by using different spectroscopic methods, molecular docking and viscometery. The competitive fluorescence and UV–Vis absorption spectroscopy indicated that the complex interacted with ctDNA via intercalating binding mode with the binding constant of 1.16?×?104 Lmol?1. The thermodynamic studies showed that the reaction between the complex and ctDNA is exothermic. Furthermore, the complex induced changes in DNA viscosity. Circular dichroism spectroscopy (CD) was employed to measure the conformational changes of ctDNA in the presence of the complex and verified intercalation binding mode. The molecular modeling results illustrated that the complex interacted via intercalation by relative binding energy of ?28.45?kJ mol?1.  相似文献   

10.
Phenyl-substituted indoloquinolines were studied for their biological activity and their DNA binding affinity. Water-soluble aminoalkyl derivatives were prepared and have shown significant in vitro anticancer activity. Unlike previous reports on the potential role of duplex DNA as target for various indoloquinoline based drugs, duplex UV melting experiments and fluorescence titrations suggest only weak and moderately strong binding of the phenyl-substituted indoloquinolines at 120 mM and 20 mM Na+ concentrations, respectively. Binding is suggested by ethidium displacement and circular dichroism experiments to be associated with drug intercalation between base pairs.  相似文献   

11.
Neutral salts activate and stabilize thermolysin. In this study, to explore the mechanism, we analyzed the interaction of 8-anilinonaphthalene 1-sulphonate (ANS) and thermolysin by ANS fluorescence. At pH 7.5, the fluorescence of ANS increased and blue-shifted with increasing concentrations (0–2.0?μM) of thermolysin, indicating that the anilinonaphthalene group of ANS binds with thermolysin through hydrophobic interaction. ANS did not alter thermolysin activity. The dissociation constants (Kd) of the complex between ANS and thermolysin was 33?±?2?μM at 0?M NaCl at pH 7.5, decreased with increasing NaCl concentrations, and reached 9?±?3?μM at 4?M NaCl. The Kd values were not varied (31?34?μM) in a pH range of 5.5?8.5. This suggests that at high NaCl concentrations, Na+ and/or Cl ions bind with thermolysin and affect the binding of ANS with thermolysin. Our results also suggest that the activation and stabilization of thermolysin by NaCl are partially brought about by the binding of Na+ and/or Cl ions with thermolysin.  相似文献   

12.
Various inhibitors were tested for their inhibitory effects on soybean urease. The Ki values for boric acid, 4-bromophenylboronic acid, butylboronic acid, and phenylboronic acid were 0.20?±?0.05?mM, 0.22?±?0.04?mM, 1.50?±?0.10?mM, and 2.00?±?0.11?mM, respectively. The inhibition was competitive type with boric acid and boronic acids. Heavy metal ions including Ag+, Hg2+, and Cu2+ showed strong inhibition on soybean urease, with the silver ion being a potent inhibitor (IC50 = 2.3?×?10?8 mM). Time-dependent inhibition studies exhibited biphasic kinetics with all heavy metal ions. Furthermore, inhibition studies with sodium salts of mineral acids (NaF, NaCl, NaNO3, and Na2SO4) showed that only F? inhibited soybean urease significantly (IC50 = 2.9?mM). Competitive type of inhibition was observed for this anion with a Ki value of 1.30?mM.  相似文献   

13.
The in vitro influence of potassium ion modulations, in the concentration range 2 mM–500 mM, on digoxin-induced inhibition of porcine cerebral cortex Na+/K+-ATPase activity was studied. The response of enzymatic activity in the presence of various K+ concentrations to digoxin was biphasic, thereby, indicating the existence of two Na+/K+-ATPase isoforms, differing in the affinity towards the tested drug. Both isoforms showed higher sensitivity to digoxin in the presence of K+ ions below 20 mM in the medium assay. The IC50 values for high/low isoforms 2.77 × 10? 6 M / 8.56 × 10? 5 M and 7.06 × 10? 7 M /1.87 × 10? 5 M were obtained in the presence of optimal (20 mM) and 2 mM K+, respectively. However, preincubation in the presence of elevated K+ concentration (50 – 500 mM) in the medium assay prior to Na+/K+-ATPase exposure to digoxin did not prevent the inhibition, i.e. IC50 values for both isoforms was the same as in the presence of the optimal K+ concentration. On the contrary, addition of 200 mM K+ into the medium assay after 10 minutes exposure of Na+/K+-ATPase to digoxin, showed a time-dependent recovery effect on the inhibited enzymatic activity. Kinetic analysis showed that digoxin inhibited Na+/K+-ATPase by reducing maximum enzymatic velocity (Vmax) and Km, implying an uncompetitive mode of interaction.  相似文献   

14.
H P Hopkins  W D Wilson 《Biopolymers》1987,26(8):1347-1355
Enthalpy changes (ΔHB) for the binding of ethidium (a monocation) and propidium (a dication) to calf thymus DNA have been determined calorimetrically in piperazine-N, N′-bis(2-ethanesulfonic acid) buffer with the fluoride ion as the counterion. Heats of dilution for the fluoride salts of ethidium and propidium were substantially less than the corresponding values found for other halide salts of these cations. At a Na+ ion concentrations of 0.019, ΔHB = ?8.3 and ?7.9 ± 0.3 kcal mol?1 for ethidium and propidium, respectively. For these two cations, just as was observed for the naphthalene monoimide (monocation) and diimide (dication) [H. P. Hopkins, K. A. Stevenson, and W. D. Wilson, (1986) J. Sol. Chem. 15 , 563–579], ΔHB is within the same experimental error for both cations. Apparently, charge–charge interactions in DNA–cation complexes produce only small changes in the enthalpy for the system. In the concentration range 0.019–0.207, the ΔHB values for propidium did not depend appreciably on the Na+ ion concentration, and a similar pattern was shown to exist for ethidium. When these results were combined with ΔGB values for the binding of these cations to DNA, we found the variation of ΔSB with Na+ ion concentration to be remarkably close to the predictions of modern polyelectrolyte theory, i.e., propidium binding to DNA causes approximately twice as many Na+ ions to be released into the bulk solution as does the binding of ethidium. The much stronger binding of propidium, relative to ethidium, at low ionic strengths is thus seen to be primarily due to entropic effects.  相似文献   

15.
Abstract

Ferulic acid (FA), a dietary phenolic acid compound, is proved to possess numerous biological activities. Hence, this study was devoted to explore the interaction between FA and calf thymus DNA (ctDNA) by UV???vis absorption, fluorescence, circular dichroism (CD) spectroscopy combined with multivariate curve resolution-alternating least-squares (MCR???ALS) and molecular docking studies. The concentration curves and the pure spectra of compositions (FA, ctDNA and FA???ctDNA complex) were obtained by MCR???ALS approach to verify and monitor the interaction of FA with ctDNA. The groove binding mode between FA and ctDNA was confirmed by the results of melting analysis, viscosity measurements, single-stranded DNA experiments, and competitive studies. The binding constant of FA???ctDNA complex was 4.87?×?104 L mol?1 at 298?K. The values of enthalpy (ΔH°) and entropy (ΔS°) changes in the interaction were ?16.24?kJ mol?1 and 35.02?J mol?1 K?1, respectively, indicating that the main binding forces were hydrogen bonds and hydrophobic interactions. The result of CD spectra suggested that a decrease in right-handed helicity of ctDNA was induced by FA and the DNA conformational transition from the B-form to the A-form. The results of docking indicated that FA binding with ctDNA in the minor groove. These findings may be conducive to understand the interaction mechanism of FA with ctDNA and the pharmacological effects of FA.

Communicated by Ramaswamy H. Sarma

  相似文献   

16.
Ultraviolet-visible (UV-vis) spectra, fluorescence spectra, electrochemistry, and the thermodynamic method were used to discuss the interaction mode between the inclusion complex of hematoxylin with β-cyclodextrin and herring sperm DNA. On the condition of physiological pH, the result showed that hematoxylin and β-cyclodextrin formed an inclusion complex with binding ratio nhematoxylin:nβ-cyclodextrin = 1:1. The interaction mode between β-cyclodextrin-hematoxylin and DNA was a mixed binding, which contained intercalation and electrostatic mode. The binding ratio between β-cyclodextrin-hematoxylin and DNA was nβ-cyclodextrin -hematoxylin:nDNA = 2:1, binding constant was K? 298.15K = 5.29 × 104 L·mol?1, and entropy worked as driven force in this action.  相似文献   

17.
The effect of magnesium ions on the parameters of the DNA helix-coil transition has been studied for the concentration range 10?6–10?1M at the ionic strengths of 10?3M Na+. Special attention has been given to the region of low ion concentrations and to the effect of polyvalent metallic impurities present in DNA. It has been shown that binding with Mg++ increases the DNA stability, the effect being observed mainly in the concentration range 10?6–10?4M. At[Mg++]>10?2M the thermal stability of DNA starts to decrease. The melting range extends to concentrations ~10?5M and then decreases to 7–8°C at the ion content of 10?3M. Asymmetry of the melting curves is observed at low ionic strengths ([Na+] = 10?3M) and [Mg++] ? 10?5M. The results, analyzed in terms of the statistical thermodynamic theory of double-stranded homopolymers melting in the presence of ligands, suggest that the effects observed might be due to the ion redistribution from denatured to native DNA. An experimental DNA–Mg++ phase diagram has been obtained which is in good agreement with the theory. It has been shown that thermal denaturation of the system may be an efficient method for determining the ion-binding constants for both native and denatured DNA.  相似文献   

18.
T Schleich  G R Gould 《Biopolymers》1974,13(2):327-337
Using the thermodynamic analysis and methodology of Hill (Biopolymers, 12 , 257 (1973)) for the treatment of optical thermal transition data the effects of various neutral salt additives on the stability and thermodynamics of the poly U–deoxyadenosine interactions that lead to the formation of triple-stranded helical polymer–monomer complexes have been studied. In order of increasing molar effectiveness as polyU–deoxyadenosine complex stability perturbants (pH 7 and in the presence of 1 M NaCl), the various ions may be ranked: SO4?2 < Cl? < Br? < ClO4?; and (CH3)4 N+ < Li+ < Rb+ ~ Na+ < K+ < (CH3 CH2)4 N+ < urea < Guan+ ~ (CH3(CH2)2)4 N+. Destabilizing neutral salt additives (e.g., NaClO4) caused a decrease in the absolute magnitude of the apparent enthalpy and entropy of binding relative to the values determined in the presence of NaCl. By contrast, stabilizing additives (e.g., Na2 SO4) had the opposite effect on these parameters. Along a melting curve the apparent differential heat of complex formation calculated for the binding of deoxyadenosine to poly U in 1 M NaCl appeared to vary linearly with θ, the extent of fractional binding. For such a linear dependence it can be shown that the integral heat (usually determined calorimetrically) equals the differential heat at θ = 0.5. Correcting the apparent differential heat calculated at θ = 0.5 for ligand activity resulted in values for the integral heat of binding of deoxyadenosine to poly U in 1 M NaCl of ?13 to ?16 kcal/mol. Binding isotherms determined in the presence of different inorganic electrolytes could be superimposed provided that different temperatures were compared. However, the additive (CH3)4NCl, which has been shown to interact preferentially with A-T rich regions of DNA (Shapiro, Stannard, and Felsenfeld, Biochemistry, 8 , 3233 (1969)) resulted in a considerably broadened binding isotherm indicating less cooperativity.  相似文献   

19.
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   

20.
Methods including spectroscopy, electronic chemistry and thermodynamics were used to study the inclusion effect between γ-cyclodextrin (CD) and vitamin K3(K3), as well as the interaction mode between herring-sperm DNA (hsDNA) and γ-CD-K3 inclusion complex. The results from ultraviolet spectroscopic method indicated that VK3 and γ-CD formed 1:1 inclusion complex, with the inclusion constant Kf = 1.02 × 104 L/mol, which is based on Benesi–Hildebrand's viewpoint. The outcomes from the probe method and Scatchard methods suggested that the interaction mode between γ-CD-K3 and DNA was a mixture mode, which included intercalation and electrostatic binding effects. The binding constants were K θ25°C = 2.16 × 104 L/mol, and Kθ37°C = 1.06 × 104 L/mol. The thermodynamic functions of the interaction between γ-CD-K3 and DNA were ΔrHmθ = ?2.74 × 104 J/mol, ΔrSmθ = 174.74 J·mol?1K?1, therefore, both ΔrHmθ (enthalpy) and ΔrSmθ (entropy) worked as driven forces in this action.  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