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.
A statistical mechanical calculation of the binding properties of DNA bis-intercalators is presented, based on the sequence-generating function method of Lifson. The effects of binding by intercalation of one or both chromophores of a bifunctional intercalating agent are examined. The secular equation for a general model that includes the effects of neighbor (nearest and non-nearest) exclusion and/or cooperativity in the binding of both singly and doubly intercalated ligands is derived. Numerical results for binding curves are presented for a more restricted model in which each type of bound ligand rigorously excludes its nearest neighbor and the total number of sites covered by a doubly intercalated ligand is variable. At low values of free ligand concentration bis-intercalation dominates the binding process, while at high value of free ligand concentration, intercalation of only one chromophore per ligand becomes significant due to the unavailability of contiguous free sites required for bis-intercalation. Also, depending on the binding parameters, the free energy of the system can be lowered by a loss of doubly intercalated ligands in favor of singly intercalated ones. Corresponding to this transition in binding mode, the average number of sites occupied by a bound ligand decreases from that characteristic of bis-intercalation to that characteristic of mono-intercalation as free ligand concentration increases. An analysis of Scatchard plots describing bis-intercalation is presented.  相似文献   

2.
Binding isotherms can provide quantitive information regarding the stability of a molecular complex. Theorectical studies in recent years have been directed to systems in which a single ligand can exclude more than one polymer site (excluded-site effect). This system has minium of thre parameters to describe the binding data: the intrinsic binding constant, B; the remote-neighbor cooperative paramaters, σq and the number of excluded sites, q. It is suggested in the present communication that precise values for these three parameters can be obtained by utilizing the characteristics of two forms of data representation: θ vs ln m and θ/m vs θ, where θ is the degree of saturation (0?θ?1) and m is the molality of free ligand. The matrix generation method is used to obtain empirical equations relating the midpoint location and slope at the midpoint of the θ vs ln m plot to the three molecular parameters. A modified Scatchard theory is also presented for highly cooperative systems, which results in an expression relaing the maximum in the θ/m vs θ plot to the molecular parameters σq and q, thus providing the third equation for the three unknown parameters. The novel method f analysis is illustrated with the AMP-poly(L -arginine) and oligocytidylate–T7 DNA sstems.  相似文献   

3.
The factors that determine the binding of a chromophore between the base pairs in DNA intercalation complexes are dissected. The electrostatic potential in the intercalation plane is calculated using an accurate ab initio based distributed multipole electrostatic model for a range of intercalation sites, involving different sequences of base pairs and relative twist angles. There will be a significant electrostatic contribution to the binding energy for chromophores with a predominantly positive electrostatic potential, but this varies significantly with sequence, and somewhat with twist angle. The usefulness of these potential maps for understanding the binding of intercalators is explored by calculating the electrostatic binding energy for 9-aminoacridine, ethidium, and daunomycin in a variety of model binding sites. The electrostatic forces play a major role in the positioning of an intercalating 9-aminoacridine and a significant stabilizing role in the binding of ethidium in its sterically constrained position, but the intercalation of daunomycin is determined by the side-chain binding. Sequence preferences are likely to be determined by a complex and subtle mixture of effects, with electrostatics being just one component. The electrostatic binding energy is also unlikely to be a major determinant of the twist angle, as its variation with angle is modest for most intercalation sites. Overall, the electrostatic potential maps give guidance on how positively charged chromophores can be chemically adapted by heteroatomic substitution to optimise their binding.  相似文献   

4.
Abstract

This research is an effort to further understand the physicochemical interaction between the novel drug molecule diethidium (2,7-diamino 9-[2,7 diamino 10-nN- phenanthridium] 10- nN- phenanthridium) and its biological receptor DNA. The ultimate goal is the elucidation of this novel class of drugs as potential pharmaceutical agents. Understanding the physico- chemical properties of this drug as well as the mechanism by which it interacts with DNA should ultimately allow the rational design of novel anti-cancer or anti-viral drugs.

A novel binding structure for the diethidium cation to B-form DNA is herein described. Molecular modeling on the complex formed between diethidium and a dodecamer of double-stranded B-form DNA, CGCGAATTCGCG, has shown that this complex is indeed fully capable of participating in the formation of a stable intercalation site. It was expected that diethidium would have a mechanism of intercalation significantly different from other classical intercalators because a) Its structure, that of two perpendicular planes, each known to have excellent intercalation properties, is novel b) The linker region length is zero c) The tilt between the two planes of the drug matches the geometry of the space available to this drug in the major groove.

We have studied the complex formed when diethidium enters the central site of the B-DNA dodecamer through the major groove. The complex forms several classes of intercalation structures, which are all stable and vary from “partially intercalate” to “fully intercalated”. Multiple minimizations show the drug to be very mobile within the intercalation site. Further, some structures show organization and concomitant stiffening of the DNA above the intercalation site, with a disorganization and disruption of the regular B-DNA structure immediately below the intercalation site. This particular phenomena may be expected to lead to significantly different physicochemical properties for the diethidium complex with respect to other known intercalators, because this sort of vectorial difference in structure above and below the site of intercalation is unknown in existing intercalators, as far as the authors are aware. In addition, we expect the mechanism of interaction between drug and DNA to be described by “direct ligand transfer”, wherein the drug is transferred from duplex DNA to duplex DNA without re-entering the solvent.1

This work is the first instance known to the authors of a novel drug entity that was deduced solely by mathematical reasoning 2 and described subsequently by computational methods. Evidence that diethidium should interact with its target site DNA differently from other known intercalators is strong.  相似文献   

5.
Abstract

Actinomycin D (ActD) is a DNA-binding antitumor antibiotic that appears to act in vivo by inhibiting RNA polymerase. The mechanism of DNA binding of ActD has attracted much attention because of its strong preference for 5′-dGpdC-3′ sequences. Binding is thought to involve intercalation of the tricyclic aromatic phenoxazone ring into a GC step, with the two equivalent cyclic pentapeptide lactone substituents lying in the minor groove and making hydrogen bond contacts with the 2-amino groups of the nearest neighbor guanines. Recent studies have indicated, however, that binding is also influenced by next-nearest neighboring bases. We have examined this higher order specificity using 7-azido-actinomycin-D as a photoaffinity probe, and DNA sequencing techniques to quantitatively monitor sites of covalent photoaddition. We found that GC doublets were strongly preferred only if the 5′- flanking base was a pyrimidine and the 3′-flanking base was not cytosine. In addition we observed a previously unreported preference for binding at a GG doublet in the sequence 5′- TGGG-3′.  相似文献   

6.
DNA intercalators that have high affinity and slow kinetics are developed for potential DNA-targeted therapeutics. Although many natural intercalators contain multiple chiral subunits, only intercalators with a single chiral unit have been quantitatively probed. Dumbbell-shaped DNA threading intercalators represent the next order of structural complexity relative to simple intercalators, and can provide significant insights into the stereoselectivity of DNA-ligand intercalation. We investigated DNA threading intercalation by binuclear ruthenium complex [μ-dppzip(phen)4Ru2]4+ (Piz). Four Piz stereoisomers are defined by the chirality of the intercalating subunit (Ru(phen)2dppz) and the distal subunit (Ru(phen)2ip), respectively, each of which can be either right-handed (Δ) or left-handed (Λ). We used optical tweezers to measure single DNA molecule elongation due to threading intercalation, revealing force-dependent DNA intercalation rates and equilibrium dissociation constants. The force spectroscopy analysis provided the zero-force DNA binding affinity, the equilibrium DNA-ligand elongation Δxeq, and the dynamic DNA structural deformations during ligand association xon and dissociation xoff. We found that Piz stereoisomers exhibit over 20-fold differences in DNA binding affinity, from a Kd of 27 ± 3 nM for (Δ,Λ)-Piz to a Kd of 622 ± 55 nM for (Λ,Δ)-Piz. The striking affinity decrease is correlated with increasing Δxeq from 0.30 ± 0.02 to 0.48 ± 0.02 nm and xon from 0.25 ± 0.01 to 0.46 ± 0.02 nm, but limited xoff changes. Notably, the affinity and threading kinetics is 10-fold enhanced for right-handed intercalating subunits, and 2- to 5-fold enhanced for left-handed distal subunits. These findings demonstrate sterically dispersed transition pathways and robust DNA structural recognition of chiral intercalators, which are critical for optimizing DNA binding affinity and kinetics.  相似文献   

7.
The interaction of daunomycin with ctDNA and six purine–pyrimidine alternating poly-deoxynucleotides has been studied using fluorometric and uv-visible absorption methods. In the explored binding range of r > 0.05, the intercalation of the drug into the DNAs proved to be anticooperative, as indicated by the pronounced upward curvature of all the Scatchard plots obtained. The experimental data have been analyzed according to the recent theory of Friedman and Manning, which describes the polyelectrolyte effects on the site binding equilibria, drug intercalation included. We found that, accounting for the polyelectrolyte effects in the neighbor site exclusion model, the experimental data were nearly equally well described, in a wide range of binding ratios, by assuming the presence of sequence specificity effects (site size = 2 base pairs, exclusion parameter n = 1) or its absence (site size = 1 base pair, n = 1.7). The relevant results are as follows: (a) Daunomycin binds to all the DNAs considered with a stoichiometry of approximately 1 drug for every two base pairs. (b) The anticooperative nature of the interaction is essentially polyelectrolytic in origin. (c) The binding affinity shown by the drug for the different sites considered decreases in the order of Gm5C > AT > AC-GT > IC > GC > AU, indicating a stabilizing effect of the —CH3 group in position 5 of the pyrimidines. (d) The extent of quenching of the intrinsic fluorescence of daunomycin in the presence of DNA is bound to the presence, at the intercalation site, of a guanine residue, since GC, Gm5C, and AC-GT sites induce a nearly total quenching, whereas AT, AU, and IC sites act only partially in this respect. The structural results obtained from the daunomycin-d[(CGTACG)]2 crystal suggest that the 2-NH2 group of guanine might be responsible for such a phenomenon. The influence of both the temperature and the ionic strength on the free energy of drug intercalation into ctDNA, poly[d(G-C)] : poly[d(G-C)], and poly[d(A-C)] : poly[d(G-T)] is examined and discussed.  相似文献   

8.
Based on steric and electrostatic considerations, the prerequisites for binding to DNA via the intercalation mechanism are proposed. Steric contour energy curves are presented to demonstrate the region inaccessible to an intercalant. They are calculated with a 6-n (n = 14) potential. This method is a soft potential analog of an excluded-volume approach. Electrostatic contours on the steric surface illustrate the relatively positive and negative regions of the binding site. The principal intercalation sites, predicted to fit into B-DNA via a tetramer-duplex unit, and the unconstrained dimer-duplex units, obtained in crystal structures, are examined. These contours illustrate the requirements of size, conformation, and net atomic charges necessary for intercalation and optimum binding. Based on the limited space available for intercalation by the presence of the backbone and the maximum base-pair separation of 8.25 Å, an Essential Metabolite Exclusion Hypothesis is presented.  相似文献   

9.
J L Leroy  M Guéron 《Biopolymers》1977,16(11):2429-2446
The binding of Mn2+ to yeast tRNAPhe at 25°C is measured by esr, and found to depend strongly on the concentration of monovalent cations, showing the importance of electrostatic effects. In low sodium (<15mM/l.), the affinity is high and the Scatchard plots are distinctly curved. In high sodium (>50mM/l.), the affinity and the curvature are reduced. In a limited range of sodium concentrations (15–30mM/l.), the folding of tRNA which is induced by the divalent ions results in cooperative binding, leading to upwards convexity of the Scatchard plot. An electrostatic model is developed, based on a single type of binding site which we take to be the phosphates, with a binding constant for Mn2+ in the range of that found for ApA, 10 l./M. We compute the change in the binding constant due to the electrostatic potential of the distant charges (other phosphates and counterions), using a single set of parameters for all sodium concentrations. The model predicts that the plots in low sodium are curved, and a good fit to the experimental results is obtained: it is therefore not legitimate or necessary to interpret these results in terms of two types of binding sites. In high salt, the model gives plots that are only slightly curved, corresponding to weaker electrostatic effects. This shows that a search for sites with a special binding mode should be done in high salt. The computed plots are in good agreement with the data, except for slight differences concerning the first bound ions, which give a possible indication in favor of special binding. Given the observation of one special site for Mg2+ at 4°C in high sodium [Stein, A. & Crothers, D. M. (1976) Biochemistry 15 , 157–160] in E. coli tRNAfMet, we have measured the binding of Mn2+ at lower temperature. At 12°C, in both yeast tRNAPhe and E. coli tRNAfMet, the plots clearly indicate special binding. A site found in high sodium is on a very different footing from the four to six so-called strong sites unduly derived from low-salt binding plots.  相似文献   

10.
The circular dichroism spectra of eleven double-stranded DNAs, five natural with known nearest neighbor frequencies and six synthetic polydimers and polytrimers, were measured from 210 to 310 nm in the absence and presence of increasing amounts of actinomycin up to saturation. Based on the fact that the circular dichroism of nucleic acids is a nearest-neighbor frequency-dependent property, matrix analysis of the problem revealed which neighbor sets were perturbed by actinomycin, presumably by intercalation of the planar moiety of the molecule. The intercalation sites can be separated into three families. The first-neighbor units GpC and CpG are very favorable binding sites for actinomycin. ApG, CpC, ApC, TpC, and TpG appear to be less attractive sites, while ApT, TpA, and ApA are unfavorable sites.  相似文献   

11.
A series of mono-1H-benzo[c,d]indol-2-one with different amine side chains and bis-1H-benzo[c,d]indol-2-one as novel family of DNA intercalators were designed and synthesized, the contributions of aromatic chromophores and amine side chains for DNA binding properties, for example, intercalation and electrostatic binding, respectively, were evaluated. Among them, A3 tailed with N,N-dimethylamino-ethyl-ethane-1,2-diamine showed selective anti-tumor activities against cell lines A549 and P388 with IC(50) 0.428microm and 1.69microm.  相似文献   

12.
Unfused tricyclic aromatic ring systems 1-6 with one or two cationic side chains have been synthesized and their interactions with DNA and synthetic polymers probed with a variety of techniques. Molecular mechanics calculations indicate that the torsional angle between ring planes in the minimum energy conformation of the tricyclic molecules can range from 0 degree to as high as 50 degrees depending on the type of rings and substituents. Viscometric titrations with linear and supercoiled DNA, linear dichroism, and NMR studies indicated that all compounds with torsional angles of approximately 20 degrees or less bind to DNA by intercalation. The more highly twisted intercalators caused significant perturbation of DNA structure. Unfused intercalators with twist angles of approximately 20 degrees have reduced binding constants, suggesting that they could not form an optimum interaction with the DNA base pairs. Unfused intercalators with twist less than 20 degrees formed strong complexes with DNA. The structures of these unfused intercalators are more analogous to typical groove-binding molecules, and an analysis of their interaction with DNA provides a better understanding of the subtle differences between intercalation and groove-binding modes for aromatic cations. The results indicate that intercalation and groove-binding modes should be viewed as two potential wells on a continuous energy surface. The results also suggest design strategies for intercalators that can optimally complement DNA base pair propeller twist or that can induce bends in DNA at the intercalation site.  相似文献   

13.
The fluorene derivative tilorone has received great attention as a DNA intercalator and has been widely recognized as an inducer of interferon. The biological activity of tilorone is known to be related to its binding mode with DNA; however, few structural and thermodynamic studies have elaborated on this issue. This paper presents two-dimensional (2-D) NMR and isothermal titration calorimetry (ITC) for the tilorone/DNA complex, coupled with circular dichroism (CD) spectroscopy and viscosity measurements. NMR investigation suggests that tilorone binds to DNA through intercalation, showing greater affinity for insertion between AT base pairs than between CG pairs. CD spectral changes were observed for T/B (tilorone/DNA base pair molar ratio) ratios greater than the stoichiometric ratio generally expected for intercalators (i.e., T/B = 0.5, according to the neighbor-exclusion principle). However, there was a clear plateau in the CD intensity between T/B < 0.35 and T/B > 0.45. From comparison with NMR and other measurements, we postulate that CD changes below the plateau should be related to the intercalation and the latter to electrostatic interactions and nonspecific bindings. ITC data showed that DeltaH < -TDeltaS < 0, which indicated that tilorone/DNA binding is enthalpy controlled. The magnitude of Kb (the binding constant) was of the same order as that of ethidium bromide. The stoichiometric number, obtained from ITC, CD, and UV data, implied a relatively smaller value (0.28-0.35) than that of the neighbor-exclusion principle. This is because side chains located in the groove disrupt further intercalation to the adjacent sites.  相似文献   

14.
Ren J  Jenkins TC  Chaires JB 《Biochemistry》2000,39(29):8439-8447
Isothermal titration calorimetry has been used to determine the binding enthalpy and heat capacity change (DeltaC(p)()) for a series of DNA intercalators, including ethidium, propidium, daunorubicin, and adriamycin. Temperature-dependent binding enthalpies were measured directly for the ligands, from which DeltaC(p)() values of -140 to -160 cal mol(-)(1) K(-)(1) were calculated. Published van't Hoff plots were reanalyzed to obtain DeltaC(p)() values of -337 to -423 cal mol(-)(1) K(-)(1) for the binding of actinomycin D to several DNA oligonucleotide duplexes with defined sequences. Heat capacity changes for DNA intercalation were found to correlate with the alterations in solvent-accessible surface area calculated from available high-resolution structural data. Multiple linear regression was used to derive the relationship DeltaC(p)() = 0. 382(+/-0.026)DeltaA(np) - 0.121(+/-0.077)DeltaA(p) cal mol(-)(1) K(-)(1), where DeltaA(np) and DeltaA(p) are the binding-induced changes in nonpolar and polar solvent-accessible surface areas (in square angstroms), respectively. The DeltaC(p)() terms were used to estimate the hydrophobic contribution to intercalative binding free energies, yielding values that ranged from -11.2 (ethidium) to -30 kcal mol(-)(1) (actinomycin D). An attempt was made to parse the observed binding free energies of ethidium and propidium into five underlying contributions. Such analysis showed that the DNA binding behavior of these simple intercalators is driven almost equally by hydrophobic effects and van der Waals contacts within the intercalation site.  相似文献   

15.
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   

16.
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   

17.
K S Schmitz 《Biopolymers》1974,13(5):1039-1053
Simulated binding curves for the reaction 2 polymer + monomer = triple-stranded complex are presented, in which loop formation and sliding degeneracy of the polymer adsorption surface are considered. Exact calculations for a polymer chain length N of 11 units suggest that configurations of two or more loops have negligible effect on the isotherm when SW > 1, where S and W are exponential weighting factors for monomer–monomer S and polymer–polymer W nearest neighbor interactions. There is a pronounced effect, however, when SW ? 1. Limiting expressions (N ? 1, but finite) for the single-loop configurations suggest these configurations are negligible at any degree of saturation θ if θ (1 ? θ)2–k N3–k ? SW, where k is defined by the weighting factor (j + 1)?k for a ring of j units. This expression suggests that single monomer-stack configurations are the only significant contributors to the grand partition function at the midpoint of the isotherm if N3–k ? SW. Furthermore, single-loop configurations are negligible below θ = 0.5 but become dominant above the isotherm midpoint when SW ~ 1 (random binding) if 2 < k < 3. For k > 3 and N → ∞, loop configurations have no effect in any region of the random binding isotherm usually analyzed experimentally (θ < 0.95). Equivalence of matrix and sequence generating methods is also demonstrated.  相似文献   

18.
Numerous studies have documented declines in plant diversity in response to habitat loss in fragmented landscapes. However, determining the mechanisms that lead to species loss is challenging using solely a correlative approach. Here we link correlative assessments of plant community composition with seed additions for a focal species to test the hypothesis that distributions of forests plants within a fragmented landscape are limited by seed dispersal. Woody plant species richness of fragments declined as fragments (n=26) became more isolated by agricultural fields. We predicted that if these isolation effects were driven by poor dispersal rather than other effects associated with habitat loss, then plants should vary in their response to isolation in relation to their seed size (i.e., stronger effects for plants with larger seeds). As predicted under this dispersal limitation hypothesis, sensitivity of bird-dispersed shrubs to isolation was related to their seed mass, with species with heavy seeds (e.g., Lindera benzoin) exhibiting stronger declines in presence across isolation gradients than species with light seeds. Seed addition experiments were performed for Lindera benzoin in two high isolation forest fragments (nearest neighbor mean distance=803 m) where Lindera was naturally absent, and two low isolation fragments (nearest neighbor mean distance=218 m) with naturally occurring Lindera populations. Seed addition and control plots (n=50 1 m2 plots per fragment) were monitored for 13 censuses over 3 years. Across all four fragments, seed additions resulted in significant increases in Lindera seedling recruitment with no differences in final seedling establishment among fragments. However, insect herbivory was higher on Lindera seedlings in high isolation compared to low isolation fragments and was negatively correlated with seedling survival over some years. Consistent with prior work, our results confirm that seed dispersal plays a significant role in affecting plant diversity in fragmented landscapes. However, results also suggest the need for a better understanding of how additional processes, such as herbivory, may be altered as habitat is lost and what effects such changes have for forest plants.  相似文献   

19.
A simple surficial peel technique using adhesive tape was developed for quantitative removal of haptobenthic diatom communities from topographically simple substrata. The method combines high removal efficiency with low peel distortion, permitting the use of spatial statistics to test whether populations are distributed in the peel randomly or form aggregated or uniform patterns. Using this technique, the microdistribution of Cocconeis placentula Ehr. on a smooth acrylic rod was examined. Using conventional nearest neighbor analyses, a clonal population of C. placentula. characterized by an indentation of the value margin, was significantly aggregated, whereas the overall C. placentula population was uniform or aggregated depending on whether the method of analysis allowed for cell size. Using refined nearest neighbor analysis, the indented population was aggregated, and the overall population was random at distances greater than cell size. The results suggest that the indented clone was weakly motile following cell division and that its directional bearing was random.  相似文献   

20.
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   

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