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.
Influence of long-range interactions between ligands bound to DNA molecule on the character of their adsorption is studied using computer modeling. For this investigation, two calculation procedures are developed. They are based upon the method of the free energy minimum and on the partition function method. The both procedures demonstrate that in the case of a strong enough attraction between all the bound ligands their binding to DNA has the character of phase transition of the first kind. There is a break in the binding curve c(c0) where c - relative concentration of bound ligands, c0 - molar concentration of free ligands. The break occurs because there is an interval of central degrees of binding (approximately 50% of the maximum c value) that are prohibited for individual DNA molecules. Such a transition might be caused by some types of DNA condensation. Attraction between the neighboring ligands only, adjacent or/and separated by double helix regions, does not cause this effect.  相似文献   

2.
Long-range interaction between all the ligands bound to DNA molecule may give rise to adsorption with the character of phase transition of the first kind (D. Y. Lando, V. B. Teif, J. Biomol. Struct & Dynam. 18, 903-911 (2000)). In this case, the binding curve, c(c(o)), is characterized by a sudden change of the relative concentration of bound ligands ((c)) at a critical concentration of free (unbound) ligands, c(o)=c(ocr), from a low c value to a high one where c(o) is molar concentration of free ligands. Such a transition might be caused by some types of DNA condensation or changes in DNA topology. For the study of the conditions necessary for adsorption with the character of phase transition, a calculation procedure based on the method of the free energy minimum is developed. The ligand size and two types of interactions between ligands adsorbed on DNA molecule are taken into consideration: long-range interaction between all the ligands bound to DNA and contact interactions between neighboring ligands. It was found that a) Stronger long-range interaction is required for longer ligands to induce phase transition that is occurred at greater c(ocr) values; b) Pure contact interaction between neighboring ligands can not itself initiate phase transition. However contact cooperativity strongly decreases the threshold value of energy of long-range interaction necessary to give rise to the transition.  相似文献   

3.
The pKa of 3,8-diamino-6-phenyl-phenanthridine (DAPP), a nonquaternary analog of ethidium bromide, has been determined spectrophotometrically as a function of sodium ion concentration both free in solution and complexed to DNA. Unwinding angle determinations with this compound were determined with Col El DNA using ethidium bromide as a standard. The unwinding angle for DAPP was 24 ± 2° relative to 26° for ethidium, and this suggests that DAPP binds in a manner quite similar to ethidium and with no significant outside bound DAPP under these experimental conditions. Isobestic behavior was obtained on spectrophotometric pH titration above pH 5 as long as the ratio of DNA-phosphate to ligand was between 100 and 300 and the DNA phosphate concentration was approximately 0.01M or greater. The loss of isosbestic behavior which occurred below pH 5 is probably due to titration of the 8 amino group of the ligand complexed to DNA. To circumvent this problem, pKa values and the extinction coefficient of the acidic species were both determined by a computer program using experimental data obtained above pH 5. The pKa of the free compound has only a minor dependence on ionic strength, while the pKa of the ligand bound to DNA in an intercalated complex depends strongly on the sodium ion concentration. The pKa of the DAPP-DNA complex is a linear function of –log[Na+] as predicted by the ion-condensation theory of polyelectrolytes. It was determined that DAPP is essentially completely bound to DNA under the conditions of these experiments by (1) determination of apparent pKa values as a function of total DNA concentration, (2) calculation of binding constants for the neutral species of DAPP, and (3) spectral analysis of the protonated and neutral species of DAPP bound to DNA relative to DAPP free in solution. These results support the ion-condensation theory; provide an independent method for measuring ψ*, the average number of counterions associated per phosphate of DNA in the intercalated conformation; and illustrate that there are no specific pH effects or absolute pKa values for ligands bound to DNA, but only ionic-strength-dependent results.  相似文献   

4.
Binding constants and binding site sizes for the interactions of the polyamines spermine (+4), spermidine (+3), and putrecine (+2) with helical DNA have been determined as a function of ionic conditions and temperature by equilibrium dialysis using 14C-labeled polyamines. In addition, competition equilibrium dialysis has been used to determine binding parameters for the divalent cations putrescine and Mg2+ from the competitive effect of these ions on the binding of spermine or spermidine. In all cases, the logarithm of the binding constant (log Kobs) varies linearly with the logarithm of the monovalent salt concentration; the slopes d log Kobs/d log[NaCl] are proportional to the valence of the ligand, and values of the extrapolated binding constants at 1M NaCl obtained from the intercepts are small (of order 1–10M?1). In those cases examined, Kobs is insensitive to temperature; the free energy of binding is predominantly entropic. Consequently, polymines as DNA-binding ligands behave analogously to the oligolysìnes investigated previously [cf. Record, Lohman & de Haseth (1976) J. Mol. Biol. 107 , 145–158; Lohman, de Haseth & Record (1980) Biochemistry 19 , 3522–3530]. The interactions of these oligocations with DNA are predominantly electrostatic and are driven by the release of thermodynamically bound electrolyte ions from the vicinity of the DNA. The extent to which these oligocations are localized at individual phosphate binding sites or delocalized on the DNA molecule is currently not known.  相似文献   

5.
The affinity of amino acid residues to nucleic acids is probed by measurements of melting temperatures tm for the helix–coil transition at various concentrations of amino acid amides. The increase of tm on addition of ligand is described by the equation tm = t*m + αlog(1+Ktcλ), where t*m is the melting temperature in the absence of ligand, cλ the ligand concentration, and Kt the “tm-onset” constant, which is analogous to an equilibrium constant. It is shown that Kt is closely related to the affinity of the ligands to the double helix, whereas the slope α mainly reflects the preference of the ligand binding to the helix versus the coil form. In the case of the amino acid amides, α is found to be virtually independent of the nature of the side chain with few exceptions, e. g., aromatic amides. The tm-onset constant, however, strongly depends on the nature of the amino acid side chain. For simple aliphatic amino acids, the relative free energy of binding decreases with increasing hydrophobic free energy, e.g., a high affinity is found for Gly-amide and a low affinity for Leu-amide. This relation is modified by functional groups like OH in Ser-amide. The helices poly[d(A-T)], ploy[d(I-C)]. and poly[d(A-C)]·poly[d(G-T)] exhibit similar affinity scales with relatively small variations. Our results demonstrate that the hydrophilic character of double helices at their surface disfavors binding of hydrophobic ligands unless special contacts can be formed. From our results we establish an affinity scale for the binding of amino acids to double helices.  相似文献   

6.
James D. McGhee 《Biopolymers》1976,15(7):1345-1375
Theoretical calculations are conducted on the helix–coil transition of DNA, in the presence of large, cooperatively binding ligands modeled after the DNA-binding proteins of current biological interest. The ligands are allowed to bind both to helx and to coil, to cover up any number of bases or base pairs in the complex, and to interact cooperatively with their nearest neighbors. The DNA is treated in the infinite homogeneous Ising model approximation, and all calculations are done by Lifson's method of sequence-generating functions. DNA melting curves are calculated by computer in order to expolore the effects on the transition of ligand size, binding constant, free activity, and ligand–ligand cooperativity. The calculations indicate that (1) at the same intrinsic free energy change per base pair of the complexes, small ligands, for purely entropic reasons, are more effective than are large ligands in shifting the DNA melting temperature; (2) the response of the DNA melting temperature to increased ligand binding constant K and/or free ligand activity L is adequately represented at high values of KL (but not at low KL) by a simple independent site model; (3) if curves are calculated with the total amount of added ligand remaining constant and the free ligand activity allowed to vary throughout the transition, biphasic melting curves can be obtained in the complete absence of ligand–ligand cooperativity. In an Appendix, the denaturation of poly[d(A-T)] in the presence of the drug, netropsin, is used to verify some features of the theory and to illustrate how the theory can be used to obtain numerical estimates of the ligand binding parameters from the experimental melting curves.  相似文献   

7.
The process of ligand binding to a cluster of membrane-associated receptors is examined theoretically. The theoretical model proposed involves the diffusion of ligands from the solution to the disc-like cluster of receptors on the surface of the spherical cell. When the ligand hits the internal part of the disc-like cluster, it begins to move laterally until it leaves the disc through its outer surface or is bound by one of the receptors inside the disc. If the ligand leaves the cluster, it returns to the solution and hits the disc again after a certain period, etc. According to our model the transition from a diffusion-limited to a reaction-limited process of binding is determined by the dimensionless parameter Dt c/a 2, where D is the lateral diffusion coefficient,t c is the characteristic time of reaction, anda is the radius of the disc-like cluster. The forward rate constantk f turns out to be a function of . Comparing the results of our calculations ofk f with some experimental data we found that agreement is achieved at high , i.e. the process of ligand binding by clustered receptors is predominantly reaction-limited.  相似文献   

8.
Surovaya  A. N.  Grokhovsky  S. L.  Burckhardt  G.  Fritzsche  H.  Zimmer  Ch.  Gursky  G. V. 《Molecular Biology》2002,36(5):726-735
The interaction of short nucleotide duplexes with bis-netropsins, in which netropsin fragments are linked tail-to-tail via cis-diammineplatinum group (Nt–Pt(NH 3 )–Nt) or aliphatic pentamethylene chain (Nt–(CH 2 ) 5 –Nt), has been studied. Both bis-netropsins have been shown to bind to DNA oligomer 5"-CCTATATCC-3" (I) as a hairpin with parallel orientation of netropsin fragments in 1:1 stoichiometry. Monodentate binding has been detected upon binding of bis-netropsins to other duplexes of sequences 5"-CCXCC-3" [where X = TTATT (II), TTAT (III), TTTTT (IV), and AATTT (V)] along with the binding of bis-netropsins as a hairpin. The formation of dimeric antiparallel motif between the halves of two bound bis-netropsin molecules has been observed in the complexes of Nt–(CH 2 ) 5 –Nt with DNA oligomers IV and V. The ratio of binding constant of bis-netropsin as a hairpin ( 2) to monodentate binding constant ( 1) has been shown to correlate with the width and/or conformational lability of DNA in the binding site. The share of bis-netropsin bound as a hairpin decreases in the order: TATAT > TTATT > TTAAT > TTTTT > AATTT, whereas the contribution of monodentate binding rises. The minimal strong binding site for Nt–Pt(NH 3 )–Nt and Nt–(CH 2 ) 5 –Nt binding as a hairpin has been found to be DNA duplex 5"-CGTATACG-3".  相似文献   

9.
Abstract

S-shaped binding curves often characterize interactions of ligands with nucleic acid molecules as analyzed by different physicochemical and biophysical techniques. S-shaped experimental binding curves are usually interpreted as indicative of the positive cooperative interactions between the bound ligand molecules. This paper demonstrates that S-shaped binding curves may occur as a result of the “mixed mode” of DNA binding by the same ligand molecule. Mixed mode of the ligand-DNA binding can occur, for example, due to 1) isomerization or dimerization of the ligands in solution or on the DNA lattice, 2) their ability to intercalate the DNA and to bind it within the minor groove in different orientations. DNA- ligand complexes are characterized by the length of the ligand binding site on the DNA lattice (so-called “multiple-contact” model). We show here that if two or more complexes with different lengths of the ligand binding sites could be produced by the same ligand, the dependence of the concentration of the complex with the shorter length of binding site on the total concentration of ligand should be S-shaped. Our theoretical model is confirmed by comparison of the calculated and experimental CD binding curves for bis-netropsin binding to poly(dA-dT) poly(dA-dT). Bis-netropsin forms two types of DNA complexes due to its ability to interact with the DNA as monomers and trimers. Experimental S-shaped bis-netropsin-DNA binding curve is shown to be in good correlation with those calculated on the basis of our theoretical model. The present work provides new insight into the analysis of ligand-DNA binding curves.  相似文献   

10.
The interaction of methylene blue (MB) with DNA has been investigated by UV absorption spectra, Fluorescence spectra and UV-melting method. Analysis of the results of the melting experiments shows that melting temperature (T m) of the complexes increases with the [total ligand]: DNA ratio (r) at two concentrations of Na+ (2?mM Na+ and 20?mM Na+) providing support for conclusion that MB is a stabilizer of DNA helix structure. By contrast, the shapes of dependences of width of transition (ΔT) on r at low and high [Na+] are different which points to the existence of different types of binding modes of MB with DNA. UV-spectroscopy experiments and fluorescence spectra indicated that the binding modes of MB with DNA depended on r. At high r (r?>?0.25), remarkable hypochromic effect with no shift of λ max in the absorption spectra of MB was observed. The fluorescence of MB was quenched which indicated that MB was bound to phosphate groups of DNA by electrostatic interaction. At low r ratios (r?<?0.2), the absorption spectra of MB upon increasing the concentration of DNA showed gradually decrease in the peak intensities with a red shift. This phenomenon is usually associated with molecular intercalation into the base stack of the ds-DNA. Using the Scatchard’s model, the complex formation constants for MB with DNA were determined: the binding constant K?≈?6.5?×?105 and binding site size n?≈?4. Obtained data are not typical for intercalation model of ligands to DNA. Moreover, comparison between these data and our early experimental results of interaction of ethidium bromide with DNA made it possible to suggest that this binding type of MB is, more probably, semi-intercalation mode (Vardevanyan et al., 2003). This conclusion is in accordance with the analysis of the model structures of MB–DNA complexes which clearly shows the importance of solvent contributions in suggested structural form (Tong et al., 2010).  相似文献   

11.
The analysis of equilibrium binding isotherms obtained by methods such as the nitrocellulose filter binding assay, which measure the fraction. θ, of DNA to which at least one protein molecule is bound, as a function of the free protein concentration (LF) require a different type of theoretical framework from that required for analysis of conventional equilibrium binding data, in which the number of moles of protein bound per mole of DNA, θc is measured as a function of LF. The theoretical framework required to analyse equilibrium binding data generated by measuring θ(LF) is developed for co-operative and non-co-operative binding of a protein to a large number of non-specific sites and to a specific sites(s) in the presence of a large number of non-specific sites on a DNA molecule. The theory is simple to apply, equations for θ(LF) being easy to derive and evaluate, and is suitable for least-squares analysis. Two examples of the application of the theory to the analysis of experimental data are provided for the specific and non-specific binding of the EcoRI restriction endonuclease to bacteriophage λ DNA, and for the specific and non-specific binding of the enzyme dihydrofolate reductase from Lactobacillus casei to pBR322 and pWDLcB1 DNA, the latter differing from the former only in a 2.9 × 103 base-pair insert containing the L. casei dihydrofolate reductase structural gene. The theoretical and experimental advantages and disadvantages of measuring θ(LF) rather than θc(LF) are discussed.  相似文献   

12.
Abstract

A theoretical modelling of the interaction of putrescine (H+ 3N—(CH2)4—(-N+H3) with DNA is carried out, introducing two new features which make the simulation of this interaction considerably more realistic. Firstly, the DNA to which putrescine is bound is fully flexible and thus able to respond to the distorting influence of the ligand. Secondly, the effect of changing the ratio of DNA base pairs per bound ligand is explicitly modelled. In this way. we have been able to confirm the experimentally known preference of putrescine binding with AT base pairs in B-DNA, but we also show, through the new features introduced, that the nature of the binding site of the ligand and the resulting impact on DNA conformation is strongly modified by the ligand binding density.  相似文献   

13.
For a better understanding of the interactions between DNA and various acridine dyes, the binding of acridine (Acr) and 10-methylacridine (MeAcr) to native and heat-denatured calf-thymus DNA was studied in the pH range between 4 and 10 by the equilibrium dialysis and spectroscopic methods. The binding between DNA and the dyes was predominantly electrostatic. The amount of bound Acr varied with pH, mixing ratio (P/D), and the DNA conformation, and reached a maximum at pH = 5.2. The amount of bound MeAcr was constant in the pH range 5–9. The apparent binding constants of these dyes were obtained at some pH, and they were found to vary with P/D for native DNA-dye complexes. The pure spectra of bound Acr and MeAcr could be unmasked. The bound Spectra were bathochromic and hypochromic relative to the spectra of free days. Acridine bound to native DNA was shown to undergo structural changes from an acridiniumlike to a neutral acridinelike form as the pH of solutions was varied. The pK value for the transition between the bound forms was evaluated to be 7.3. The extrinsic Cotton effects of the bound dyes were observed in the DNA-Acr and-MeAcr complexes and varied with pH and the conformation of DNA.  相似文献   

14.
When increasing concentrations of methylmercuric hydroxide are added to a Cs2SO4 solution of native DNA, the buoyant density of DNA is unaltered until a critical concentration is reached above which there is a cooperative transition to denatured DNA which now binds so much CH3HgOH that it becomes very dense and nonbuoyant. As increasing concentrations of methylmercuric hydroxide are added to a Cs2So4 solution of denatured DNA, the buoyant density gradually increases, indicating a gradual increase in the amount of methylmercury cation bound. The denatured DNA methylmercury complex becomes nonbuoyant at the same concentration of methylmercuric hydroxide as does the native DNA. These results support our previous interpretation that CH3HgOH reacts with the imino NH bonds of thymine and guanine in nucleic acids. The reaction occurs more or less independently at the different binding sites for denatured DNA, but it occurs cooperatively with simultaneous denaturation for native DNA. The nature of the transition of denatured DNA to the nonbuoyant state is not known, but it is probably due to an abrupt decrease in the degree of hydration of the DNA when its density and hydrophobic character are sufficiently increased by the binding of the methylmercury cation. Direct measurements of the amount of methylmercury bound by DNA, as observed by preparative ultracentrifugation, confirm approximately the buoyant density results as to the amount of methylmercury bound. The possibility of using methylmercuric hydroxide as a reagent for the separation of complementary strands, depending on then thymine of their thymine plus guanine content, is discussed.  相似文献   

15.
Teif  V. B.  Lando  D. Yu. 《Molecular Biology》2001,35(1):106-107
A method for calculating the curves of DNA transition from linear to condensed state upon binding of condensing ligands has been developed. The character of the transition and ligand concentration necessary for condensation have been shown to be governed by the length of DNA molecule, energy and stoichiometry parameters of the DNA–ligand complex (equilibrium constant between linear and condensed form in the absence of ligands, constants for ligand binding to linear and condensed forms, the number of base pairs covered by one ligand, etc.). The results of the calculations indicate that a slight difference in the free energies of these DNA states (less than 6 cal/mol(bp) for a DNA of 500 bp) is sufficient for the existence of a stable linear state in the absence of ligands (in free DNA) and the formation of stable condensed state upon complexation.  相似文献   

16.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  相似文献   

17.
Abstract

It is argued that the enhancement of the apparent DNA contour length by the specifically binding non-intercalating drug netropsin (Nt) (Reinert et al., NAR 9,2335, 1981) at very low Nt/DNA-phosphate ratios essentially is the result of an abolition of periodically arranged intrinsic helix bends in A · T rich tracts of base pairs.

In the preceding paper the existence of pronounced DNA tertiary structure components has been postulated for (two species of) natural eukaryotic DNA. The resulting model suggests local apparent solenoid-related DNA tertiary structure components at high sodium ion concentration cs, partly/totally molten out at 45/60 C. With decreasing cs the tertiary structure components have been found to be gradually reduced, at least below cs = 0.010 M, as titration viscometrically revealed by a gradual rise of the apparent DNA contour length (Reinert et al., JBSD 9, 537, 1991).

Hence, we performed titration viscometric analyses about Nt interaction with calf thymus DNA (ctDNA) at cs = 0.075 M, 0.010 M and 0.004 M Na+. The concomitant DNA conformational changes are quantitatively described in terms of the relative changes of both DNA persistence length and hydrodynamically operative apparent DNA contour length for the three first resolved interaction modes below a Nt/DNA-P ratio of 0.03.

These experiments, together with previous respective analyses at cs = 0.20 M Na+ and different temperatures (I.e.), suggest that those DNA sites binding Nt most strongly predominantly are responsible for the formation of solenoid-related DNA tertiary structure components. Most probably these are A tract-containing sequences. As the essential factor for their apparent elongation effect at low Na+ concentrations, a gradual alteration of the number of base pairs per helix turn seems to occur below cs = 0.010 M Na+ and, concomitantly, a change in phasing between intrinsic helix bends and helix screw.  相似文献   

18.
A method for calculating the curves of DNA transition from linear to condensed state upon binding of condensing ligands has been developed. The character of the transition and ligand concentration necessary for condensation have been shown to be governed by the length of DNA molecule, energy and stoichiometry parameters of DNA-ligand complex (equilibrium constant between linear and condensed form in the absence of ligands, constants for ligand binding to linear and condensed forms, the number of base pairs covered by one ligand, etc.). The results of the calculations indicate that only slight difference in the free energies of these states in free DNA (less than 6 cal/mole(bp) for DNA of 500 bp long) is sufficient for the existence of stable linear state in the absence of ligands (in free DNA) and the formation of stable condensed state upon complexation.  相似文献   

19.
Three minor-groove binding ligands have been used to study the characteristics of two d(GA·CT)n DNAs embedded in longer DNA fragments. The binding of mithramycin, netropsin or Thia-Net to these sequences has been studied using DNAse I footprinting. None of these ligands appeared to bind to d(GA·CT)5 nor to d(GA·CT)22 extensively, although with mithramycin some protected bonds were detected at the very edge of these sequences. In general, these small ligands did not enhance the DNAse I cleavage patterns at the alternating d(GA·CT)n flanking sequences located near DNA regions where the drug was bound. The d(GA·CT)n sequences could act as a rigid block in which it is not easy to propagate structural changes, whereas other sequences flanking the binding sites showed cleavage enhancements.  相似文献   

20.
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