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.
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   

2.
3.
The interaction of synthetic polynucleotide double strands with a natural lipid, oleic acid, was examined in diluted aqueous solutions by circular dichroism spectra, UV-absorption measurements, and surface plasmon resonance biosensor investigations. The investigations were performed with defined double and triple stranded oligo- and polydeoxyribonucleotides. Whereas duplexes are influenced by oleic acid ligandation, which could not be removed by ethanol dialysis procedure, no binding occurs to triple stranded DNA. The spectroscopic results indicate that oleic acid shows molecular recognition to AT b.p. motifs by groove binding. GC tracts - in particular alternating d[G-C] motifs - are strongly influenced by ligand interaction up to a ratio of one molecule per two base pairs. Likewise, the spectroscopic and morphologic changes in the supramolecular association of the complexes after treatment occur even after dialysis procedure. This was monitored with scanning force microscopy (SFM) as well. Additionally, monolayers of biotinylated DNA duplexes were immobilized on a streptavidin sensor-layer for surface plasmon resonance (SPR) observations. Small portions of the ligand were injected in continuous flow. Loosely bound molecules were removed by washing procedure. Injections of sodium hydroxide denature the DNA, releasing the tightly bound effectors. The amount of tightly bound oleic acid molecules was determined at one molecule per 2-3 base pairs. As consequence, a new mechanism of regulation of gene expression at nuclear membrane or by lipids inside DNA double helix has to be discussed.  相似文献   

4.
Landt SG  Tipton AR  Frankel AD 《Biochemistry》2005,44(17):6547-6558
The local geometry of a DNA helix can influence protein recognition, but the sequence-specific features that contribute to helix structure are not fully understood, and even less is known about how RNA helix geometry may affect protein recognition. To begin to understand how local or global helix structure may influence binding in an RNA model system, we generated a series of DNA analogues of HIV and BIV TAR RNAs in which ribose sugars were systematically substituted in and around the known binding sites for argininamide and a BIV Tat arginine-rich peptide, respectively, and measured their corresponding binding affinities. For each TAR interaction, binding occurs in the RNA major groove with high specificity, whereas binding to the all-DNA analogue is weak and nonspecific. Relatively few substitutions are needed to convert either DNA analogue of TAR into a high-affinity binder, with the ribose requirements being restricted largely to regions that directly contact the ligand. Substitutions at individual positions show up to 70-fold differences in binding affinity, even at adjacent base pairs, while two base pairs at the core of the BIV Tat peptide-RNA interface are largely unaffected by deoxyribose substitution. These results suggest that the helix geometries and unique conformational features required for binding are established locally and are relatively insulated from effects more than one base pair away. It seems plausible that arginine-rich peptides are able to adapt to a mosaic helical architecture in which segments as small as single base steps may be considered as modular recognition units.  相似文献   

5.
Metal-mediated base pairs formed by the interaction between metal ions and artificial bases in oligonucleotides have been developed for potential applications in nanotechnology. We recently found that a natural C:C mismatched base pair bound to an Ag(+) ion to generate a novel metal-mediated base pair in duplex DNA. Preparation of the novel C-Ag-C base pair involving natural bases is more convenient than that of metal-mediated base pairs involving artificial bases because time-consuming base synthesis is not required. Here, we examined the thermodynamic properties of the binding between the Ag(+) ion and each of single and double C:C mismatched base pair in duplex DNA by isothermal titration calorimetry. The Ag(+) ion specifically bound to the C:C mismatched base pair at a 1:1 molar ratio with 10(6) M(-1) binding constant, which was significantly larger than those for nonspecific metal ion-DNA interactions. The specific binding between the Ag(+) ion and the single C:C mismatched base pair was mainly driven by the positive dehydration entropy change and the negative binding enthalpy change. In the interaction between the Ag(+) ion and each of the consecutive and interrupted double C:C mismatched base pairs, stoichiometric binding at a 1:1 molar ratio was achieved in each step of the first and second Ag(+) binding. The binding affinity for the second Ag(+) binding was similar to that for the first Ag(+) binding. Stoichiometric binding without interference and negative cooperativity may be favorable for aligning multiple Ag(+) ions in duplex DNA for applications of the metal-mediated base pairs in nanotechnology.  相似文献   

6.
High-resolution NMR techniques (proton and 19F) have been used to study the interactions between several DNA oligonucleotides with varying length of AT base pairs and the synthetic pyrrole-containing compound (P1-F4S-P1), which has properties similar to the DNA minor groove binding drug distamycin A. When this two-fold symmetrical DNA binding molecule is added to the self-complementary DNA oligomers, the resulting complex exhibits an NMR spectrum without any doubling of individual resonances, consistent with a two-fold symmetry of the complex. This is in contrast to all other complexes studied so far. The minimum length of an AT stretch for specific ligand binding is judged to be greater than 4 base pairs. Inter-molecular proton nuclear Overhauser effects between the ligand molecule and a DNA dodecamer d(CGCAAATTTGCG) provide evidence that P1-F4S-P1 binds DNA in the minor groove and interacts with the middle AT base pairs. The presence of a specific interaction between P1-F4S-P1 and DNA is conclusively demonstrated by 19F NMR studies, in which four previously chemically equivalent fluorine nuclei in the free molecule become two non-equivalent pairs (yielding an AB quartet pattern) upon the binding of P1-F4S-P1 to DNA duplex. A sequence-dependent binding behavior of P1-F4S-P1 is evident by comparing the 19F NMR spectra of the complexes between P1-F4S-P1 and two different but related DNA dodecamers, d(CGCAAATTTGCG) and d(CGCTTTAAAGCG). P1-F4S-P1 binds more strongly to the former dodecamer with an association constant of approximately 1 X 10(3) M-1.  相似文献   

7.
Tandem helical repeats have emerged as an important DNA binding architecture. DNA glycosylase AlkD, which excises N3- and N7-alkylated nucleobases, uses repeating helical motifs to bind duplex DNA and to selectively pause at non-Watson–Crick base pairs. Remodeling of the DNA backbone promotes nucleotide flipping of the lesion and the complementary base into the solvent and toward the protein surface, respectively. The important features of this new DNA binding architecture that allow AlkD to distinguish between damaged and normal DNA without contacting the lesion are poorly understood. Here, we show through extensive mutational analysis that DNA binding and N3-methyladenine (3mA) and N7-methylguanine (7mG) excision are dependent upon each residue lining the DNA binding interface. Disrupting electrostatic or hydrophobic interactions with the DNA backbone substantially reduced binding affinity and catalytic activity. These results demonstrate that residues seemingly only involved in general DNA binding are important for catalytic activity and imply that base excision is driven by binding energy provided by the entire substrate interface of this novel DNA binding architecture.  相似文献   

8.
Visualization of double stranded DNA in gels with the binding of the fluorescent dye ethidium bromide has been a basic experimental technique in any molecular biology laboratory for >40 years. The interaction between ethidium and double stranded DNA has been observed to be an intercalation between base pairs with strong experimental evidence. This presents a unique opportunity for computational chemistry and biomolecular simulation techniques to benchmark and assess their models in order to see if the theory can reproduce experiments and ultimately provide new insights. We present molecular dynamics simulations of the interaction of ethidium with two different double stranded DNA models. The first model system is the classic sequence d(CGCGAATTCGCG)2 also known as the Drew–Dickerson dodecamer. We found that the ethidium ligand binds mainly stacked on, or intercalated between, the terminal base pairs of the DNA with little to no interaction with the inner base pairs. As the intercalation at the terminal CpG steps is relatively rapid, the resultant DNA unwinding, rigidification, and increased stability of the internal base pair steps inhibits further intercalation. In order to reduce these interactions and to provide a larger groove space, a second 18-mer DNA duplex system with the sequence d(GCATGAACGAACGAACGC) was tested. We computed molecular dynamics simulations for 20 independent replicas with this sequence, each with ∼27 μs of sampling time. Results show several spontaneous intercalation and base-pair eversion events that are consistent with experimental observations. The present work suggests that extended MD simulations with modern DNA force fields and optimized simulation codes are allowing the ability to reproduce unbiased intercalation events that we were not able to previously reach due to limits in computing power and the lack of extensively tested force fields and analysis tools.  相似文献   

9.
4', 6-Diamidine-2-phenylindole forms fluorescent complexes with synthetic DNA duplexes containing AT, AU and IC base pairs; no fluorescent complexes were observed with duplexes containing GC base pairs or with duplexes containing a single AT base pair sandwiched between GC pairs. The binding site size is one molecule of dye per 3 base pairs. The intrinsic binding constants are higher for alternating sequence duplexes than for the corresponding homopolymer pairs. With the exception of the four-stranded helical poly rI which exhibits considerable fluorescence enhancement upon binding of the ligand, none of the single- or multi- stranded polyribonucleotides and ribo-deoxyribonucleotide hybrid structures form fluorescent complexes with the dye. Poly rI is the only RNA which forms a DNA B-like structure (Arnott et al. (1974) Biochem. J. 141, 537). The B conformation of the helix and the absence of guanine appear to be the major determinants of the specificity of the fluorescent binding mode of the dye. Nonfluorescent interactions of the dye with polynucleotides are nonspecific; UV absorption and circular dichroic spectra demonstrate binding to synthetic single- and double-stranded DNA and RNA analogs, including those containing GC base pairs.  相似文献   

10.
Gene regulation by steroids is tightly coupled to hormone concentration and stereochemistry. A key step is binding of hormones to receptors which interact with consensus DNA sequences known as hormone response elements (HREs). The specificity and strength of hormone binding do not correlate well with hormonal activity suggesting an additional step involving recognition of ligand by the gene. Stereospecific fit of hormones between base pairs and correlation of fit with hormonal activity led to the proposal that such recognition involves insertion of hormone into DNA. Here, the feasibility of insertion was investigated using computer models of the glucocorticoid receptor DNA binding domain bound to its HRE. The site reported to accommodate glucocorticoids was found in the HRE and was exposed to permit unwinding at this locus. The resulting cavity in the unwound DNA/receptor interface fit cortisol remarkably well; cortisol formed hydrogen bonds to both the receptor and DNA. Current experimental evidence is generally consistent with ligand binding domains of receptors undergoing a conformational change which facilitates transfer of the ligand into the unwound DNA/receptor interface. We propose this step is rate limiting and alterations in receptor, DNA or hormone which attenuate insertion impair hormonal regulation of gene function.  相似文献   

11.
We examined in detail the DNA interaction of the nuclear receptors NGFI-B and steroidogenic factor 1 (SF-1) by using a series of gain-of-function domain swaps. NGFI-B bound with high affinity as a monomer to a nearly linear DNA molecule. The prototypic zinc modules interacted with a half-site of the estrogen receptor class, and a distinct protein motif carboxy terminal to the zinc modules (the A box) interacted with two A/T base pairs 5' to the half-site. SF-1 bound in the same manner as NGFI-B, with an overlapping but distinct sequence requirement 5' to the half-site. The key features that distinguished the NGFI-B and SF-1 interactions were an amino group in the minor groove of the SF-1 binding sequence and an asparagine in the SF-1 A box. These results define a common mechanism of NGFI-B and SF-1 DNA binding, which may underlie a competitive mechanism of gene regulation in steroidogenic tissues that express these proteins. This monomer-DNA interaction represents a third paradigm of DNA binding by nuclear receptors in addition to direct and inverted dimerization.  相似文献   

12.
Ornithine decarboxylase (ODC) is a pyridoxal-5'-phosphate-dependent (PLP) enzyme that catalyzes the biosynthesis of the polyamine putrescine. Similar to other PLP-dependent enzymes, an active site Lys residue forms a Schiff base with PLP in the absence of substrate. The mechanistic role of this residue (Lys-69) in catalysis by Trypanosoma brucei ODC has been studied by analysis of the mutant enzymes, in which Lys-69 has been replaced by Arg (K69R ODC) and Ala (K69A ODC). Analysis of K69A ODC demonstrated that the enzyme copurified with amines (e.g. putrescine) that were tightly bound to the active site through a Schiff base with PLP. In contrast, on the basis of an absorption spectrum of K69R ODC, PLP is likely to be bound to this mutant enzyme in the aldehyde form. Pre-steady-state kinetic analysis of the reaction of K69R ODC with L-Orn and putrescine demonstrated that the rates of both the product release (k(off.Put) = 0.0041 s(-)(1)) and the decarboxylation (k(decarb) = 0.016 s(-)(1)) steps were decreased by10(4)-fold in comparison to wild-type ODC. Further, the rates of Schiff base formation between K69R ODC and either substrate or product have decreased by at least 10(3)-fold. Product release remains as the dominant rate-limiting step in the reaction (the steady-state parameters for K69R ODC are k(cat) = 0.0031 s(-)(1) and K(m) = 0.18 mM). The effect of mutating Lys-69 on the decarboxylation step suggests that Lys-69 may play a role in the proper positioning of the alpha-carboxylate for efficient decarboxylation. K69R ODC binds diamines and amino acids with higher affinity than the wild-type enzyme; however, Lys-69 does not mediate substrate specificity. Wild-type and K69R ODC have similar ligand specificity preferring to bind putrescine over longer and shorter diamines. Kinetic analysis of the binding of a series of diamines and amino acids to K69R ODC suggests that noncovalent interactions in the active site of K69R ODC promote selective ligand binding during Schiff base formation.  相似文献   

13.
We report dichroism and equilibrium binding studies of netropsin (Net) and distamycin A3 (Dist) binding to deoxyribonucleic acid (DNA). We show that at low degrees of binding (r) to calf thymus DNA, Net induces a considerable increase in the apparent DNA length (14 A/drug molecule bound), closely analogous to the results reported earlier for Dist. In addition, we show that chicken erythrocyte DNA shows length changes similar to those of calf thymus DNA upon distamycin binding. DNA length reaches a maximum at 1 bound drug/20-30 base pairs and then decreases to its initial value by r = 0.1. This effect is not seen for two other DNAs with nearly identical A + T base pair content and may therefore arise from the details of base sequence or base modification in eukaryotic DNA. We also show that Dist binding to calf thymus DNA at low r values is positively cooperative and shows a DNA affinity which is primarily nonionic. We demonstrate that independent of the DNA to which they are bound, the Net and Dist transition moments are inclined by 43 +/- 3 degrees from the helix axis, consistent with the idea that both drugs bind inside and parallel to the DNA small groove. From dichroism measurements, we show that the conformational change induced in calf thymus DNA by Dist does not kink or bend the helix and does not substantially alter the average inclination of the bases. Finally, we outline a statistical mechanical theory for calculation of binding isotherms when binding is coupled to a DNA structural change.  相似文献   

14.
Abstract

We prepared extracts of Alisma orientalis from Sichuan and Fujian Province, China. Based on the ratio of alisol B 23-acetate (23B) to alisol A 24-acetate (24A) in two Alisma orientalis extracts, we prepared two mixtures of 24A and 23B (24A:23B?=?1:3 or 1:10). The antitumor molecular mechanism of the monomers 24A and 23B, the two mixtures and the effective components of Alisma orientalis from different habitats were studied. The MTT assay suggested that the difference in the antitumor activity of Alisma orientalis from different habitats was correlated to the ratio of 24A to 23B. The multi-spectroscopic analysis suggested that the effective components, the monomers and mixtures interacted with c-myc DNA in a partial intercalation manner. The binding strength of the alisol acetates to c-myc DNA was consistent with the anticancer activity, indicating that c-myc DNA was the anticancer target. The molecular simulation indicated that the mixtures were all directly bound to different base pairs of c-myc DNA for a superimposed effect, which led to the binding strength of the mixtures to c-myc DNA was stronger than that of the monomers. The molecules in the 1:3 mixture were all bound to different base pairs of c-myc DNA. However, for the 1:10 mixture, seven molecules of 23B bound to the side chain of 24A, resulting in the mixture with a long chain structure which increased the steric hindrance of 24A. As a result, affinity between 24A and c-myc DNA in the 1:10 mixture was weaker than that in the 1:3 mixture.

The antitumor molecular mechanism of the alisol monomers 24A and 23B, the mixtures with different proportions and the effective components of Alisma orientalis from different habitats were studied. The order of the antitumor activity was as follows: Sichuan?>?Fujian, 24A-23B (1:3) > 24A-23B (1:10) > 23B?>?24A. The antitumor activity of Alisma orientalis from different habitats was consistent with the mixtures which were designed according to the contents of the active ingredients of the medicinal materials, indicating that the antitumor activity of Alisma orientalis from Sichuan is better than that from Fujian which is related to the contents of 24A and 23B and the proportion of 1:3 is better than 1:10. The binding strength of the mixtures to c-myc DNA was consistent with the anticancer activity. The mixtures were all directly bound to different base pairs of c-myc DNA for a superimposed effect, which led to the strength of the interaction of the mixtures to c-myc DNA was stronger than that of the monomers. For the 24A-23B (1:3) mixture, the four small molecules bound to c-myc DNA directly and interacted with different base pairs of c-myc DNA. While for the 24A-23B (1:10) mixture, 24A and three 23B molecules interacted with c-myc DNA, the remaining seven 23B molecules bound to the side chain of 24A, which increased the steric hindrance. The binding of the mixture to c-myc DNA was decreased.

Communicated by Ramaswamy H. Sarma  相似文献   

15.
M W Van Dyke  P B Dervan 《Biochemistry》1983,22(10):2373-2377
The DNA binding sites for the antitumor, antiviral, antibiotics chromomycin, mithramycin, and olivomycin on 70 base pairs of heterogeneous DNA have been determined by using the (methidiumpropyl-EDTA)iron(II) [MPE x Fe(II)] DNA cleavage inhibition pattern technique. Two DNA restriction fragments 117 and 168 base pairs in length containing the lactose operon promoter-operator region were prepared with complementary strands labeled with 32P at the 3' end. MPE x Fe(II) was allowed to partially cleave the restriction fragment preequilibrated with either chromomycin, mithramycin, or olivomycin in the presence of Mg2+. The preferred binding sites for chromomycin, mithramycin, and olivomycin in the presence of Mg2+ appear to be a minimum of 3 base pairs in size containing at least 2 contiguous dG x dC base pairs. Many binding sites are similar for the three antibiotics; chromomycin and olivomycin binding sites are nearly identical. The number of sites protected from MPE x Fe(II) cleavage increases as the concentration of drug is raised. For chromomycin/Mg2+, the preferred sites on the 70 base pairs of DNA examined are (in decreasing affinity) 3'-GGG, CGA greater than CCG, GCC greater than CGA, CCT greater than CTG-5'. The sequence 3'-CGA-5' has different affinities, indicating the importance of either flanking sequences or a nearly bound drug.  相似文献   

16.
We have studied the DNA sequence binding preference of the antitumour antibiotic nogalamycin by DNase-I footprinting using a variety of DNA fragments. The DNA fragments were obtained by cloning synthetic oligonucleotides into longer DNA fragments and were designed to contain isolated ligand-binding sites surrounded by repetitive sequences such as (A)n.(T)n and (AT)n. Within regions of (A)n.(T)n, clear footprints are observed with low concentrations of nogalamycin (< 5 microM), with apparent binding affinities for tetranucleotide sequences which decrease in the order TGCA > AGCT = ACGT > TCGA. In contrast, within regions of (AT)n, the ligand binds best to AGCT; binding to TCGA and TGCA is no stronger than to alternating AT. Within (ATT)n, the preference is for ACGT > TCGA. Although each of these binding sites contains all four base pairs, there is no apparent consensus sequence, suggesting that the selectivity is affected by local DNA dynamic and structural effects. At higher drug concentrations (> 25 microM), nogalamycin prevents DNAse-I cleavage of (AT)n but shows no interaction with regions of (AC)n.(GT)n. Regions of (A)n.(T)n, which are poorly cut by DNase I, show enhanced rates of cleavage in the presence of low concentrations of nogalamycin, but are protected from cleavage at higher concentrations. We suggest that this arises because drug binding to adjacent regions distorts the DNA to a structure which is more readily cut by the enzyme and which is better able to bind further ligand molecules.  相似文献   

17.
We have developed a technique of partially-restrained molecular mechanics enthalpy minimisation which enables the sequence-dependence of the DNA binding of a non-intercalating ligand to be studied for arbitrary sequences of considerable length (greater than = 60 base-pairs). The technique has been applied to analyse the binding of berenil to the minor groove of a 60 base-pair sequence derived from the tyrT promoter; the results are compared with those obtained by DNAse I and hydroxyl radical footprinting on the same sequence. The calculated and experimentally observed patterns of binding are in good agreement. Analysis of the modelling data highlights the importance of DNA flexibility in ligand binding. Further, the electrostatic component of the interaction tends to favour binding to AT-rich regions, whilst the van der Waals interaction energy term favours GC-rich ones. The results also suggest that an important contribution to the observed preference for binding in AT-rich regions arises from lower DNA perturbation energies and is not accompanied by reduced DNA structural perturbations in such sequences. It is therefore concluded that those modes of DNA distortion favourable to binding are probably more flexible in AT-rich regions. The structure of the modelled DNA sequence has also been analysed in terms of helical parameters. For the DNA energy-minimised in the absence of berenil, certain helical parameters show marked sequence-dependence. For example, purine-pyrimidine (R-Y) base pairs show a consistent positive buckle whereas this feature is consistently negative for Y-R pairs. Further, CG steps show lower than average values of slide while GC steps show lower than average values of rise. Similar analysis of the modelling data from the calculations including berenil highlights the importance of DNA flexibility in ligand binding. We observe that the binding of berenil induces characteristic responses in different helical parameters for the base-pairs around the binding site. For example, buckle and tilt tend to become more negative to the 5'-side of the binding site and more positive to the 3'-side, while the base steps at either side of the centre of the site show increased twist and decreased roll.  相似文献   

18.
The interaction of Hoechst 33258 with DNA has been examined to help clarify the mechanisms of banding. 1. In agreement with previous studies Hoechst fluorescence is enhanced to a greater degree in AT-rich compared to GC-rich DNA. 2. Hoechst causes an increase in the DNA Tm which is greater at the higher AT content of the DNA. 3. There is a decrease in extinction coefficient and shift in the adsorption spectra to a higher wavelength when Hoechst binds to DNA. 4. DNA is completely precipitated at a ratio of one dye molecular per base pair, and this precipitation is not affected by salt. 5. There is no increase in viscosity or change in the circular dichroism of DNA when bound to Hoechst. These findings suggest Hoechst does not bind to DNA by intercalation or by ionic interaction with the phosphate groups, but rather binds by an attachment to the outside of the double DNA helix by interacting with the base pairs. This type of binding allows greater sensitivity to the base composition than occurs with intercalating agents. In this respect its binding is similar to that of dibutyl proflavine (Muller et al., 1973).  相似文献   

19.
Sequence-specific high mobility group (HMG) box factors bind and bend DNA via interactions in the minor groove. Three-dimensional NMR analyses have provided the structural basis for this interaction. The cognate HMG domain DNA motif is generally believed to span 6-8 bases. However, alignment of promoter elements controlled by the yeast genes ste11 and Rox1 has indicated strict conservation of a larger DNA motif. By site selection, we identify a highly specific 12-base pair motif for Ste11, AGAACAAAGAAA. Similarly, we show that Tcf1, MatMc, and Sox4 bind unique, highly specific DNA motifs of 12, 12, and 10 base pairs, respectively. Footprinting with a deletion mutant of Ste11 reveals a novel interaction between the 3' base pairs of the extended DNA motif and amino acids C-terminal to the HMG domain. The sequence-specific interaction of Ste11 with these 3' base pairs contributes significantly to binding and bending of the DNA motif.  相似文献   

20.
Interaction of netropsin, distamycin A and a number of bis-netropsins with DNA fragments of definite nucleotide sequence was studied by footprinting technique. The nuclease protection experiments were made at fixed DNA concentration and varying ligand concentrations. The affinity of ligand for a DNA site was estimated from measurements of ligand concentration that causes 50% protection of the DNA site. Distribution pattern of the protected and unprotected regions along the DNA fragment was compared with the theoretically expected arrangement of the ligand along the same DNA. The comparison led us to the following conclusions: 1. Footprinting experiments show that at high levels of binding the arrangement of netropsin molecules along the DNA corresponds closely to the distribution pattern expected from theoretical calculations based on the known geometry of netropsin--DNA complex. However, the observed differences in the affinity of netropsin for various DNA sequences is markedly greater than that expected from theoretical calculations. 2. Netropsin exhibits a greater selectivity of binding than that expected for a ligand with three specific reaction centers associated with the antibiotic amide groups. It binds preferentially to DNA regions containing four or more successive AT pairs. Among 13 putative binding sites for netropsin with four or more successive AT pairs there are 11 strong binding sites and two weaker sites which are occupied at 2 D/P less than or equal to 1/9 and 2 D/P = 1/4, respectively. 3. The extent of specificity manifested by distamycin A is comparable to that shown by netropsin although the molecule of distamycin A contains four rather than three amide groups. At high levels of binding distamycin A occupies the same binding sites on DNA as netropsin does. 4. The binding specificity of bis-netropsins is greater than that of netropsin. Bis-netropsins can bind to DNA in such a way that the two netropsin-like fragments are implicated in specific interaction with DNA base pairs. However, the apparent affinity of bis-netropsins estimated from footprinting experiments is comparable with that of netropsin for the same DNA region. 5. At high levels of binding bis-netropsins and distamycin A (but not netropsin) can occupy any potential site on DNA irrespectively of the DNA sequence. 6. Complex formation with netropsin increases sensitivity to DNase I at certain DNA sites along with the protection effect observed at neighboring sites.  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