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.
F A Tanious  S F Yen  W D Wilson 《Biochemistry》1991,30(7):1813-1819
The interaction of a symmetric naphthalene diimide with alkylamino substituents at each imide position was investigated with the alternating sequence polymers, poly[d(A-T)]2 and poly[d(G-C)]2. Spectrophotometric binding studies indicate strong binding of the diimide to both sequences although the GC binding constant is 20-25 times larger than the AT binding constant. Analysis of the effects of salt concentration on the binding equilibria shows that the diimide forms two ion pairs in its complex with both polymers as expected for a simple dication. Stopped-flow kinetics experiments demonstrate that the diimide both associates and dissociates from DNA more slowly than classical intercalators with similar binding constants. Analysis of salt concentration effects on dissociation kinetics rate constants (kd) reveals that slopes in log kd versus log [Na+] plots are only approximately half the value obtained for classical dicationic intercalators that have both charged groups in the same groove. These kinetics results support a threading intercalation model, with one charged diimide substituent in each of the DNA grooves rather than with both side chains in the same groove, for the diimide complex with DNA. In the rate-determining step of the mechanism for dissociation of a threading complex only one ion pair is broken; the free side chain can then slide between base pairs to put both diimide side chains in the same groove, and this is followed by rapid full dissociation of the diimide. This sequential release of ion pairs makes the dissociation slope for dicationic threading intercalators more similar to the slope for classical monocationic intercalating ligands.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Binding constants Kobs, expressed per site and evaluated in the limit of zero binding density, are quantified as functions of salt (sodium acetate) concentration for the interactions of the oligopeptide ligand KWK6NH2 (designated L8+, with ZL = 8 charges) with three single-stranded DNA oligomers (ss dT-mers, with |ZD| = 15, 39, and 69 charges). These results provide the first systematic experimental information about the effect of changing |ZD| on the strength and salt dependence of oligocation-oligonucleotide binding interactions. In a comparative study of L8+ binding to poly dT and to a short dT oligomer (|ZD| = 10),. Proc. Natl. Acad. Sci. USA. 93:2511-2516) demonstrated the profound thermodynamic effects of phosphate charges that flank isolated nonspecific L8+ binding sites on DNA. Here we find that both Kobs and the magnitude of its power dependence on salt activity (|SaKobs|) increase monotonically with increasing |ZD|. The dependences of Kobs and SaKobs on |ZD| are interpreted by introducing a simple two-state thermodynamic model for Coulombic end effects, which accounts for our finding that when L8+ binds to sufficiently long dT-mers, both DeltaGobso = -RT ln Kobs and SaKobs approach the values characteristic of binding to poly-dT as linear functions of the reciprocal of the number of potential oligocation binding sites on the DNA lattice. Analysis of our L8+-dT-mer binding data in terms of this model indicates that the axial range of the Coulombic end effect for ss DNA extends over approximately 10 phosphate charges. We conclude that Coulombic interactions cause an oligocation (with ZL < |ZD|) to bind preferentially to interior rather than terminal binding sites on oligoanionic or polyanionic DNA, and we quantify the strong increase of this preference with decreasing salt concentration. Coulombic end effects must be considered when oligonucleotides are used as models for polyanionic DNA in thermodynamic studies of the binding of charged ligands, including proteins.  相似文献   

3.
The binding of the antitumor agents SN-16814 nd SN-13232 to various DNA's in solution was monitored by CD and UV absorption measurements. In addition comparative studies with dA.dT containing duplex DNA of the related ligands SN-6136 and SN-6324 were included with respect to effects of structural variations. In general all four ligands show a dA.dT preference in their binding affinity to DNA. Differences were observed for the reaction of SN-16814 which contains bicyclic ring system: it has a lower base pair selectivity, shows some affinity to poly(dG-dC).poly(dG-dC), poly(rA).poly(rU) and poly(rU). The binding mechanism of SN-16814 is associated with a significant time dependent binding effect in CD spectra and UV absorption in case of reaction with poly(dA).poly(dT) and poly(dI).poly(dC) indicating a slow kinetics. The preferred binding to dA.dT base pairs in DNA decreases in the order from SN-61367 greater than SN-13232 greater than SN-6324,SN-16814 as judged from CD titration studies, salt dissociation and melting temperature data. Competitive binding experiments with netropsin (Nt) or distamycin-5 revealed that SN-16814 and SN-13232 are displaced from poly(dA.dT).poly(dA-dT) suggesting that both ligands are less strongly bound than Nt and Dst-5 within the minor groove of B-DNA. These studies are consistent with results of the DNAse I cleavage of poly(dA-dT).poly(dA-dT) which show the same relative order of inhibition of the cleavage reaction due to ligand binding. The results suggest that the variability of the DNA binding and dA.dT sequence specificity may reside in the adaptability of benzamide-type ligands in the helical groove which is influenced by distinct structural modifications of the ligand conformation.  相似文献   

4.
S-shaped binding curves often characterize interactions of ligands with nucleic acid molecules as analyzed by different physico-chemical 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   

5.
Optical methods, such as fluorescence, circular dichroism and linear flow dichroism, were used to study the binding to DNA of four symmetrical cyanine dyes, each consisting of two identical quinoline, benzthiazole, indole, or benzoxazole fragments connected by a trimethine bridge. The ligands were shown to form a monomer type complex into the DNA minor groove. The complex of quinoline-containing ligand with calf thymus DNA appeared to be the most resistant to ionic strength, and it did not dissociate completely even in 1 M NaCl. Binding of cyanine dyes to DNA could also be characterized by possibility to form ligand dimers into the DNA minor groove, by slight preference of binding to AT pairs, as well as by possible intercalation between base pairs of poly(dG)-poly(dC). The correlation found between the binding constants to DNA and the extent of cyanine dyes hydrophobicity estimated as the n-octanol/water partition coefficient is indicative of a significant role of hydrophobic interactions for the ligand binding into the DNA minor groove.  相似文献   

6.
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   

7.
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   

8.
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   

9.
Abstract

The binding of the antitumor agents SN-16814 nd SN-13232 to various DNA's in solution was monitored by CD and UV absorption measurements. In addition comparative studies with dA · dT containing duplex DNA of the related ligands SN-6136 and SN-6324 were included with respect to effects of structural variations. In general all four ligands show a dA · dT preference in their binding affinity to DNA.

Differences were observed for the reaction of SN-16814 which contains bicyclic ring system: it has a lower base pair selectivity, shows some affinity to poly(dG-dC) · poly(dG-dC), poly(rA) · poly(rU) and poly(rU). The binding mechanism of SN-16814 is associated with a significant time dependent binding effect in CD spectra and UV absorption in case of reaction with poly(dA) · poly(dT) and poly(dI) · poly(dC) indicating a slow kinetics.

The preferred binding to dA · dT base pairs in DNA decreases in the order from SN-61367 > SN-13232 > SN-6324, SN-16814 as judged from CD titration studies, salt dissociation and melting temperature data. Competitive binding experiments with netropsin (Nt) or distamycin-5 revealed that SN-16814 and SN-13232 are displaced from poly(dA-dT) · poly(dA-dT) suggesting that both ligands are less strongly bound than Nt and Dst-5 within the minor groove of B-DNA. These studies are consistent with results of the DNAase I cleavage of poly(dA-dT) · poly(dA-dT) which show the same relative order of inhibition of the cleavage reaction due to ligand binding. The results suggest that the variability of the DNAbinding and dA · dT sequence specificity may reside in the adaptability of benzamide-type ligands in the helical groove which is influenced by distinct structural modifications of the ligand conformation.  相似文献   

10.
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   

11.
Many examples exist of charged amino acids that play a role in attracting or holding a charged ligand toward or inside an oppositely charged binding pocket of the protein. For example, the enzymes superoxide dismutase, triose-phosphate isomerase, and acetylcholinesterase can steer ligands toward their oppositely charged binding pockets or gorges. Interestingly, in our Brownian dynamics simulations of a phosphate-binding protein, we discovered that negatively charged phosphate (HPO(2-)(4)) could make its way into the negatively charged binding pocket. In fact, the phosphate-binding protein exhibits counterintuitive kinetics of association. That is, one would expect that the rate of association would increase on increases to the ionic strength since the interaction between the ligand, with a charge of -2, and the electronegative binding pocket would be repulsive and greater screening should reduce this repulsion and increase the rate of association. However, the opposite is seen-i.e., the rate of association decreases on increases in the ionic strength. We used Brownian dynamics techniques to compute the diffusion limited association rate constants between the negatively charged phosphate ligand and several open forms of PBP (wild-type and several mutants based on an x-ray structure of open-form PBP, mutant T141D). With the appropriate choices of reaction criteria and molecular parameters, the ligand was able to diffuse into the binding pocket. A number of residues influence binding of the ligand within the pocket via hydrogen bonds or salt bridges. Arg135 partially neutralizes the charges on the HPO(2-)(4) ligand in the binding pocket, allowing it to enter. It is also found that the positive electrostatic patches above and below the binding entrance of PBP contribute the major attractive forces that direct the ligand toward the surface of the protein near the binding site.  相似文献   

12.
We have measured the thermal melting profile for poly[d(AT)].poly[d(TA)] as a function of concentration of three trivalent cations: spermidine, me8spermidine, and hexammine cobalt(III). Using McGhee's (1976) theory of DNA melting in the presence of ligands, we have estimated association constants Kh, Kc and binding site sizes nh, nc for binding to double-helical (h) and single-stranded (c) polynucleotide. The results are as follows: (table; see text) The binding parameters for spermidine and hexammine cobalt(III) to double helical molecules agree fairly well with direct equilibrium dialysis measurements, and are in reasonable accord with predictions of counterion condensation theory. However, despite their identical charges, the three ligands bind to single-stranded DNA with quite different affinities. Estimates of the charge spacing of single-stranded DNA suggest that poly[d(AT)] is less elongated in the presence of spermidine and hexammine cobalt(III) than it is when complexed with me8spermidine.  相似文献   

13.
The equilibrium association constants for the binding of a wide variety of effecting ligands of the lac repressor were measured by equilibrium dialysis. Also, detailed investigations of the apparent rate of dissociation of repressor-operator comples as a function of ligand concentration were carried out for several inducers and anti-inducers. The affinity of repressor-ligand comples for operator DNA was evaluated from the specific rate constants at saturating concentrations of effecting ligand. By fitting the experimental data depicting the functional dependence of the rate of dissociation upon ligand concentrations to calculated curves, assuming simple models of the induction mechanism, the equilibrium association constant for the binding of effecting ligand to repressor-operator comples was determined. Inducers reduce the affinity of lac repressor for operator DNA by a factor of approximately 1000 under standard conditions; the extent of destabilization depends on Mg2+ ion concentration. Anti-inducers increase the affinity of repressor for operator at most a factor of five. Only one neutral ligand, which binds to repressor without altering the stability of repressor-operator comples, was found. No homotropic or heterotropic interactions in the binding of effecting ligands either to repressor or to repressor-operator complex are evident.  相似文献   

14.
Hargreaves VV  Schleif RF 《Biochemistry》2008,47(13):4119-4128
The equilibrium dissociation constant of the DNA binding domain of interferon regulatory factor 1 (IRF1 DBD) for its DNA binding site depends strongly on salt concentration and salt type. These dependencies are consistent with IRF1 DBD binding to DNA, resulting in the release of cations from the DNA and both release of anions from the protein and uptake of a cation by the protein. We demonstrated this by utilizing the fact that the release of fluoride from protein upon complex formation does not contribute to the salt concentration dependence of binding and by studying mutants in which charged residues in IRF1 DBD that form salt bridges with DNA phosphates are changed to alanine. The salt concentration dependencies of the dissociation constants of wild-type IRF1 DBD and the mutants R64A, D73A, K75A, and D73A/K75A were measured in buffer containing NaF, NaCl, or NaBr. The salt concentration and type dependencies of the mutants relative to wild-type IRF1 DBD provide evidence of charge neutralization by solution ions for R64 and by a salt bridge between D73 and K75 in buffer containing chloride or bromide salts. These data also allowed us to determine the number, type, and localization of condensed ions around both IRF1 DBD and its DNA binding site.  相似文献   

15.
The interaction between the B-form specific ligands netropsin (Nt) and distamycin-3 (Dst-3) and DNA duplexes has been studied under conditions of salt concentration and low water activity that modify the polymer conformation into a non-B DNA form, putatively a Z-like form. Three polymers with strict alternating purine-pyrimidine sequences and GC content from 100-0% have been tested: poly(dG-dC) . poly(dG-dC), poly(dA-dC) . poly(dG-dT) and poly(dA-dT) . poly(dA-dT). The titrations by Nt and Dst-3 were followed by circular dichroism. Although specific binding of Nt to the Z-form of poly(dG-dC) . poly(dG-dC) does not occur, Nt reverses this Z structure to the B-type conformation; Dst-3 is, however, totally inefficient. The presumed non-B or Z-like structure of poly(dA-dC) . poly(dG-dT) is reversed to the B-form upon interaction with Nt; Dst-3 also induces this reversal but at higher ligand ratios. The modified B-structure of poly(dA-dT) . poly(dA-dT) in low water activity is efficiently reversed to the B-form by interaction with both Nt and Dst-3.  相似文献   

16.
A theory of polyelectrolyte effects on site-binding equilibria is generalized to multivalent ligands, multivalent supporting salt, intercalation, and multiple-site exclusion. The theory, which contains no adjustable parameters, except the number of sites excluded by a bound ligand, gives the dependence of the equilibrium constant on the binding fraction and the salt concentration. The theory is compared with prior experimental data for the dissociation of poly(acrylic acid), the binding of magnesium to polyphosphate, and the binding of ethidium and actionomycin D to DNA. The theory predicts the binding fraction dependence of the dissociation constant of poly(acrylic acid) well. The theory predicts the binding fraction dependence of the association constant of the binding of Mg2+ to polyphosphate well, if either one or two phosphates are bound by a magnesium ion. We conclude that polyelectrolyte effects on drug-DNA equilibria must be substantial. It follows that an incorrect estimate of the number of sites excluded by a bound drug molecule (because of its size or some other nonpolyelectrolyte effect) can be obtained from binding data if polyelectrolyte effects are ignored. The estimate is also within the context of, and subject to the validity of, the model used to describe the nonpolyelectrolyte contribution to binding. Our results suggest that, subject to these conditions, the anticooperativity of the binding of ethidium to DNA might be explained solely in terms of polyelectrolyte effects, and without reference to multiple-site exclusion, if sequence-specificity effects can be safely ignored. Our results also suggest that as few as two base pairs might be excluded by an actinomycin molecule. The theory gives fairly good agreement for the salt-concentration dependence of the association constant of all of the systems studied, including the complex of the neutral drug actinomycin with DNA.  相似文献   

17.
Electrospray ionization mass spectrometry (ESI-MS) has been used to determine the dissociation constants (K(D)s) and binding stoichiometry for tobramycin and paromomycin with a 27-nucleotide RNA construct representing the A-site of the 16S ribosomal RNA. K(D) values determined by holding the ligand concentration fixed are compared with K(D) values derived by holding the RNA target concentration fixed. Additionally, the effect of solution conditions such as the amount of organic solvent present and the amount of salt present in the solution on the K(D) measurement is investigated. It is shown that the preferred method for determining dissociation constants using ESI-MS is holding the RNA target concentration fixed below the expected K(D) and titrating the ligand. K(D) measurements should also be carried out at as high as possible salt concentration to minimize nonspecific binding due primarily to electrostatic interactions. For tobramycin, two nonequivalent binding sites were found with K(D1) = 352 nM and K(D2) = 9 microM. For paromomycin, there is only one binding site with K(D) = 52 nM.  相似文献   

18.
In studies of interactions of ligands with nucleic acid molecules, S-shaped curves are often obtained. In this work we show that the sigmoidal shape of the curves is due to a competition of ligand molecules for DNA binding sites. If a ligand that binds to DNA can form two types of complexes with the binding constants K1 and K2 and binding site sizes L1 and L2, respectively, then in the event that K1 < K2 and L1 < L2, the dependence of the concentration of complexes of the first type on the concentration of the ligand added to solution can be S-shaped.  相似文献   

19.
A few drug-like molecules have recently been found to bind poly(A) and induce a stable secondary structure (Tm ≈ 60°C), even though this RNA homopolymer is single-stranded in the absence of a ligand. Here, we report results from experiments specifically designed to explore the association of small molecules with poly(A). We demonstrate that coralyne, the first small molecule discovered to bind poly(dA), binds with unexpectedly high affinity (Ka >107 M−1), and that the crescent shape of coralyne appears necessary for poly(A) binding. We also show that the binding of similar ligands to poly(A) can be highly cooperative. For one particular ligand, at least six ligand molecules are required to stabilize the poly(A) self-structure at room temperature. This highly cooperative binding produces very sharp transitions between unstructured and structured poly(A) as a function of ligand concentration. Given the fact that junctions between Watson–Crick and A·A duplexes are tolerated, we propose that poly(A) sequence elements and appropriate ligands could be used to reversibly drive transitions in DNA and RNA-based molecular structures by simply diluting/concentrating a sample about the poly(A)-ligand ‘critical concentration’. The ligands described here may also find biological or medicinal applications, owing to the 3′-polyadenylation of mRNA in living cells.  相似文献   

20.
Although peptide nucleic acids (PNAs) are neutral by themselves, they are usually appended with positively charged lysine residues to increase their solubility and binding affinity for nucleic acid targets. Thus obtained cationic PNAs very effectively interact with the designated duplex DNA targets in a sequence-specific manner forming strand-invasion complexes. We report on the study of the nonspecific effects in the kinetics of formation of sequence-specific PNA-DNA complexes. We find that in a typical range of salt concentrations used when working with strand-invading PNAs (10-20 mM NaCl) the PNA binding rates essentially do not depend on the presence of nontarget DNA in the reaction mixture. However, at lower salt concentrations (<10 mM NaCl), the rates of PNA binding to DNA targets are significantly slowed down by the excess of unrelated DNA. This effect of nontarget DNA arises from depleting the concentration of free PNA capable of interacting with DNA target due to adhesion of positively charged PNA molecules on the negatively charged DNA duplex. As expected, the nonspecific electrostatic effects are more pronounced for more charged PNAs. We propose a simple model quantitatively describing all major features of the observed phenomenon. This understanding is important for design of and manipulation with the DNA-binding polycationic ligands in general and PNA-based drugs in particular.  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