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.
The inhibition of ribonucleic acid polymerase by acridines   总被引:8,自引:6,他引:2       下载免费PDF全文
1. The aminoacridines, proflavine (3,6-diaminoacridine) and 9-aminoacridine, and a hydrogenated derivative, 9-amino-1,2,3,4-tetrahydroacridine, were shown to inhibit in vitro the DNA-primed RNA polymerase of Escherichia coli. The inhibition is strong with both proflavine and 9-aminoacridine, but weak with 9-amino-1,2,3,4-tetrahydroacridine. 2. The extent to which the three acridines bind to calf-thymus DNA in the enzyme medium was studied spectrophotometrically. The extent of binding decreases in the order: proflavine, 9-aminoacridine, 9-amino-1,2,3,4-tetrahydroacridine. Some evidence was also obtained for interaction between the nucleoside triphosphate substrates and proflavine or 9-aminoacridine; no such interaction was detectable with 9-amino-1,2,3,4-tetrahydroacridine. 3. Although the amount of acridine bound to DNA increases with increasing inhibition, a stage is reached where an increase in acridine concentration still causes an increase in inhibition, with practically no increase in the amount bound to DNA. 4. Plots of reciprocal rates against the reciprocal of DNA concentration were linear and had a common intercept when proflavine or 9-aminoacridine was present. Similar relations were obtained when the reciprocal concentration of nucleoside triphosphates was plotted. The observations are interpreted kinetically in terms of a competitive inhibition of the enzyme by proflavine or 9-aminoacridine and of a kinetic role for the DNA analogous to ;activation'. 5. This suggests that inhibitory acridine molecules can occupy the sites on the RNA polymerase that are specific for binding the nucleoside triphosphate substrate or the bases of the DNA, when these become accessible during the copying process.  相似文献   

2.
The effect of sequence on the binding of 9-aminoacridine to DNA has been investigated by studying its interaction with deoxydinucleoside phosphates of different sequences using proton nuclear magnetic resonance. Quantitative binding information can be obtained by comparison of the proton chemical shift behavior of 9-aminoacridine upon addition of dinucleoside phosphate to various models for the interaction using least-squares computer fitting procedures. The simplest model that fits the data includes (1) dimerization of 9-aminoacridine and (2) a mixture of 1:1 and 2:1 (dinucleoside phosphate/9-aminoacridine) complexes. The computed parameters allow comparison of binding constants and stereochemistry for different sequences. The 1:1 complexes seem to involve interaction of the ring nitrogen with the backbone phosphate and stacking of one or both chromophores on the acridine; preference in binding is observed for alternating (purine-pyrimidine or pyrimidine-purine) over non-alternating (purine-purine) dinucleoside phosphates. The 2:1 complexes involve intercalation of the acridine between two complementary dinucleoside phosphate strands with weak sequence preferences in binding. The stereochemistry of intercalation differs between non-alternating purine-purine sequences and the alternating pyrimidine-purine or purine-pyrimidine sequences in having the 9-aminoacridine stacked with the purines of one strand rather than straddling the purines on opposite strands. The difference in stereochemistry could possibly be a determining factor in frameshift sequence specificity.  相似文献   

3.
The rigidity of the linking chain of bifunctional intercalators in the ditercalinium series was shown to be critical for antitumor activity. In order to study the influence of the rigidity of the linking chain on the DNA-binding properties of DNA bifunctional intercalators, fluorescent 9-aminoacridine and 2-methoxy-6-chloro-9-aminoacridine analogues with chains of variable rigidity were synthesized. 1H-NMR studies show that the conformation of 9-aminoacridine dimers is almost independent of the nature of the linking chain. A strong self-stacking of the aromatic rings of the 2-methoxy-6-chloro-9-aminoacridine is observed for dimers with flexible chains but not for those with rigid chains. All the dimers having a linking chain long enough to bisintercalate in DNA according to the excluded site model are indeed bisintercalators. The kinetic association constant of all monomers and dimers for poly[d(A-T)].poly[d(A-T)] are in the same range (2-4 x 10(7) M-1 s-1). The large increase of DNA binding affinity observed for the dimers is always associated with the expected decrease of the dissociation rate constant. The effect of chain rigidity and pH on the calf thymus DNA binding of 9-aminoacridine and 2-methoxy-6-chloro-9-aminoacridine dimers is quite different. In the series of 9-aminoacridine the pKa of the dimers remains high and therefore no difference of DNA-binding affinity is observed between pH 5 and 7.4. The rigidity of the linking chain does not significantly alter the DNA-binding affinity. In the 2-methoxy-6-chloro-9-aminoacridine series, the pKa of all dimers became smaller than the physiological pH and a dramatic decrease of DNA-binding affinity is observed when the pH is increased from pH 5 to 7.4. This decrease appears significantly smaller for dimers with rigid chains. A similar dramatic decrease of binding affinity at pH 7.4 is not observed for poly[d(A-T)].poly[d(A-T)]. This factor makes these dimers strongly specific for the alternating polymer at pH 7.4.  相似文献   

4.
Interaction of a spin-labeled 9-aminoacridine with DNA was studied by electron spin resonance spectroscopy. Accurate determination of the binding parameters, equilibrium dissociation constant (KD), and total number of ligand-binding sites was obtained using Scatchard and Lineweaver-Burk plots. The competition between 9-aminoacridine and its spin-labeled derivative was examined by a similar analysis of the spin-label signals. The practical interest of this method lies in the fact that the precision and the simplicity of the measurements allow the quantitative determination of the binding capacity of any intercalative drug which interacts specifically with adenine/thymine bases of DNA.  相似文献   

5.
E L Loechler  J King 《Biochemistry》1986,25(20):5858-5864
Acridine dyes bound to the condensed DNA within phage particles sensitize them to inactivation by visible light. The mechanism involves absorption of photons by an acridine/DNA complex, generating singlet oxygen, which covalently damages nearby proteins needed for DNA injection [Bryant, J., & King, J. (1985) J. Mol. Biol. 180, 837-863]. Acridines and related dyes interact with double-stranded DNA through a number of binding modes. To determine in condensed phage DNA the binding mode responsible for this inactivation, we have studied the formation of the DNA/acridine target complexes for photoinactivation. Analysis of the kinetics of 9-aminoacridine binding to Salmonella phage P22 particles revealed the formation of two binding species, one of which appeared more rapidly and was apparently an intermediate in the formation of the second. The rapidly forming species represented DNA sites with intercalated acridines, while the more slowly forming species represented the subsequent binding of additional acridine molecules to the DNA backbone of sites already containing intercalated dye. The rates of photoinactivation correlated with the rate of binding of 9-aminoacridine to the DNA backbone. This suggests that the most effective species for sensitizing phage to light-induced damage has acridine molecules stacked alongside the backbone of a region with intercalated molecules.  相似文献   

6.
This study investigated the effect of CpG methylation on the DNA binding of cisplatin analogues with an attached aminoacridine intercalator. DNA-targeted 9-aminoacridine carboxamide Pt complexes are known to bind at 5′-CpG sequences. Their binding to methylated and non-methylated 5′-CpG sequences was determined and compared with cisplatin. The damage profiles of each platinum compound were quantified via a polymerase stop assay with fluorescently labelled primers and capillary electrophoresis. Methylation at 5′-CpG was shown to significantly increase the binding intensity for the 9-aminoacridine carboxamide compounds, whereas no significant increase was found for cisplatin. 5′-CpG methylation had the largest effect on the 9-ethanolamine-acridine carboxamide Pt complex, followed by the 9-aminoacridine carboxamide Pt complex and the 7-fluoro complex. The methylation state of a cell’s genome is important in maintaining normal gene expression, and is often aberrantly altered in cancer cells. An analogue of cisplatin which differentially targets methylated DNA may be able to improve its therapeutic activity, or alter its range of targets and evade the chemoresistance which hampers cisplatin efficacy in clinical use.  相似文献   

7.
M D Topal 《Biochemistry》1984,23(11):2367-2372
9-Aminoacridine was found to inhibit dNTP incorporation into DNA homopolymer duplexes by phage T4 DNA polymerase in vitro. Systematic variation of the molar ratio of 9-aminoacridine to DNA, to DNA polymerase, and to DNA precursors demonstrated that this inhibition at 9-aminoacridine concentrations below 10 microM was mainly due to interaction of 9-aminoacridine with the DNA and suggested that the basis for the preferential inhibition of incorrect precursor incorporation was destabilization of the DNA growing point. Consistent with destabilization, 9-aminoacridine stimulated the hydrolysis of correctly base paired DNA by the 3'-5' exonuclease activity of phage T4 DNA polymerase. This is the first indication to my knowledge that an intercalating dye destabilizes the DNA growing point, whereas it raises the overall Tm of the DNA. At 9-aminoacridine concentrations above 10 microM overall incorporation of dNTPs was inhibited by 9-aminoacridine interaction with the DNA polymerase. A possible explanation for the induction of both deletion and addition frameshift mutations by 9-aminoacridine during DNA biosynthesis is discussed in light of growing-point destabilization.  相似文献   

8.
In this study, the detailed DNA sequence specificity of four acridine Pt complexes was examined and compared with that of cisplatin. The DNA sequence specificity was determined in a telomere-containing DNA sequence using a polymerase stop assay, with a fluorescent primer and an automated capillary DNA sequencer. The Pt compounds included an acridine intercalating moiety that was modified to give a 9-aminoacridine derivative, a 7-methoxy-9-aminoacridine derivative, a 7-fluoro-9-aminoacridine derivative and a 9-ethanolamine-acridine derivative. Compared with cisplatin, the DNA sequence specificity was most altered for the 7-methoxy-9-aminoacridine compound, followed by the 9-aminoacridine derivative, the 7-fluoro-9-aminoacridine compound and the 9-ethanolamine-acridine derivative. The DNA sequence selectivity for the four acridine Pt complexes was shifted away from runs of consecutive guanines towards single guanine bases, especially 5′-GA dinucleotides and sequences that contained 5′-CG. The sequence specificity was examined in telomeric and non-telomeric DNA sequences. Although it was found that telomeric DNA sequences were extensively damaged by the four acridine Pt complexes, there was no extra preference for telomeric sequences.  相似文献   

9.
The ability of 9-aminoacridine to induce mutagenic lesions during DNA replication in vitro was investigated. The ampicillinase gene of pBR322 was replicated in vitro in the presence of 9-aminoacridine. Transfection of the replicated DNA into Escherichia coli gave Amps mutants. Determination of the base changes in 76 of these mutants indicated that the spectrum of mutations induced by 9-aminoacridine was consistent with its action in vivo. Both large (407-base) and small (1- and 2-base) deletions were induced at repetitive sequences. The frequency of deletion mutations depended on the identity of the base deleted and sequences surrounding the deletions. The characteristics of the frameshift mutations induced were consistent with the interactions of 9-aminoacridine with DNA. These results establish that 9-aminoacridine can induce frameshift mutations during the replication process and provide an in vitro model of frameshift induction for mechanistic studies.  相似文献   

10.
Most of the intracellular T4 DNA made in the presence of 9-aminoacridine is of lower molecular weight than mature T4 DNA and does not get packaged into phage particles. Using a T4 DNA transformation assay, we have examined this intracellular T4 DNA for its content of 9-aminoacridine-induced revertants of certain rII gene frameshift mutations. The proportion of acridine-induced revertants in the intracellular DNA population is close to that found in the phage progency made in the presence of 9-aminoacridine. Thus, the generation of low molecular weight T4 DNA in the presence of 9-aminoacridine is not, in itself, also a mutagenic process.  相似文献   

11.
Abstract

DNA-drug complexes are important because of their pharmacological interest but, in addition, they provide a useful model to study the essential aspects of DNA recognition processes. In order to investigate the influence of ligand binding on the dynamic properties of DNA we have carried out normal mode analysis for complexes with drugs of two types: a typical intercalator, 9-aminoacridine, and a typical groove binder, netropsin. Normal modes are analysed in terms of helicoidal parameter variations with special attention being paid to global deformations of the double helix. The results show that the influence of these two drugs is very different. Intercalation of 9-aminoacridine leads to an increase in the flexibility of the intercalated dinucleotide step, with notably larger vibrational amplitudes for both roll and twist parameters compared to free DNA. In contrast, the groove binding of netropsin induces a stiffening of the DNA segment which is in contact with the drug reflected by decreased vibrational amplitudes for backbone angles and inter base pair helicoidal parameters and an increase in vibrations for adjacent base pairs in terms of buckle and propeller twist.  相似文献   

12.
We have quantitatively examined the unwinding angles for the complexes of a related series of acridine and quinoline derivatives with DNA. Ethidium bromide was used as a control for determining superhelix densities at different ionic strengths. Relative to ethidium, 9-aminoacridine and quinacrine had an essentially constant unwinding angle of approximately 17 degrees at all ionic strengths tested. The apparent unwinding angle for chloroquine and 9-amino-1,2,3,4-tetrahydroacridine was found to be ionic strength dependent, increasing with increasing ionic strength. This suggests that competitive nonintercalative binding at low ionic strengths causes an apparent lowering of the quinoline unwinding angle. This can also explain why 4-aminoquinaldine, examined at low ionic strength, gives a quite low apparent unwinding angle. Quinacrine along with chloroquinine and 9-aminoacridine approaches a limiting value for their unwinding angle of approximately 17 degrees. 4-aminoquinaldine and 9-amino-1,2,3,4-tetrahydroacridine could not be examined at an ionic strength above 0.03 because of their very low equilibrium binding constants.  相似文献   

13.
The equilibrium binding of the antitumor agent m-AMSA (4'-(9-acridinylamino) methane-sulfon-m-ansidide) has been examined by optical methods. These studies which have focused on the low bound drug concentrations (r values less than 0.02, base pairs) reveal m-AMSA to bind calf thymus DNA in a highly cooperative manner as indicated by the initial positive slope of the Scatchard plot. In contrast, the studies on the parent 9-aminoacridine under identical conditions demonstrate that this compound binds DNA in a noncooperative (neighbor exclusion) manner. The positive cooperative binding phenomenon of m-AMSA is probed as a function of ionic concentration and shown to exist over the range of salt concentrations examined (0.01 to 0.1 M); however, the magnitude of the cooperative binding is altered. This observation of cooperativity is consistent with earlier studies on biologically active compounds and may be related to such binding parameters as binding sequence selectivity and/or structural perturbations to the DNA structure.  相似文献   

14.
The long-wavelength ultraviolet (lambda approximately 420 nm) radiation induced reaction between 6-azido-2-methoxy-9-acridinylamines and supercoiled plasmid DNA results in single strand scissions and formation of covalent adducts (ratio approximately 1:10). By treating azidoacridine-photomodified DNA with piperidine at 90 degrees C, additional strand scissions are observed in a complex sequence dependent manner with an overall preference for T greater than or equal to G greater than C much greater than A. The resulting DNA fragments migrate as 5'-phosphates in polyacrylamide gels. Photofootprinting of the binding site of RNA-polymerase on promoter DNA is demonstrated with an azido-9-acridinylamino-octamethylene-9-aminoacridine. Similar experiments using 9-amino-6-azido-2-methoxyacridine indicate that this reagent recognizes changes in the DNA conformation induced by RNA polymerase binding, in relation to open complex formation.  相似文献   

15.
The DNA binding characteristics of a mono-, di- and trimeric derivative of 9-aminoacridine were studied. The length of the linking carboxamidoalkyl chains was selected to allow bis- or tris-intercalation according to the excluded-site model. Measurements of DNA unwinding angle using closed circular DNA showed that the trimeric derivative behaves as a tris-intercalating agent. Nevertheless the increase of DNA binding affinity on going from dimer to trimer was found to be relatively small. This is probably related to the large structural constraint for DNA binding of the trimeric derivative. The nature of the linking chain for the design of high-affinity DNA poly-intercalating agents appears therefore critical.  相似文献   

16.
We describe the synthesis, DNA binding measurements and pharmacological properties of a series of new heterodimeric molecules, in which a 2,6-diaminopurine is linked to a 9-aminoacridine chromophore. The linking chain contains a central N,N'-disubstituted guanidine, connected to the two chromophores by polymethylenic units of variable length.  相似文献   

17.
D O Jordan  L N Sansom 《Biopolymers》1971,10(2):399-410
The influence of temperature on the binding of 9-aminoacridine and of proflavine to E. coli DNA in 10?3M NaCl solution has been determined by a spectrophotometric technique. The inadequacy of the expression normally used for the determination of the extent of binding is discussed with reference to measurements at temperatures above which dissociation of the double helix occurs. A method of determining the relative extents of binding to native and denatured DNA at elevated temperatures is described.  相似文献   

18.
D G Dalgleish  A R Peacocke 《Biopolymers》1971,10(10):1853-1863
The circular dicrosim (CD) spectra of complexes of DNA with ethidiun bromnide, profiavine, 9-aminoacridine and 4-etliyl-9-amino-acridine have been determined between 220 and 450 nm, the range lieing extended to 600 nm for ethidiufm bromide. The variation of the magnitude of the visible and near—ultraviolet CD spectra of ethidium bromide—DNA complexes with the amount of ligand bound (r) suggests a common binding position with profiavine. On the other hand, 4-ethyl-9-aminoacndine complexed to DNA shows CD spectra not distinguishable from those of 9-aminnoacnidmc in both the visible and ultraviolet. The interpretation of these results with respect to the stereochemistry of the DNA-ligand complexes is discussed.  相似文献   

19.
Two 1-nitro-9-aminoacridine dimers were prepared: one bearing a spermine flexible linking chain, compound 4, the other a rigid dipiperidine-type linker, compound 7. Both dimers elicited a higher affinity constant for DNA than the parent monomeric drug nitracrine 2. This affinity was several orders lower than what was found for other dimeric compounds having the same linkers and no nitro group on the acridine ring (3, 5, 6 and 8). Bisintercalation was evidenced for compound 4 by viscosimetric measurements. In the absence of dithiothreitol, an inhibitory effect of RNA synthesis in vitro was observed for all the tested compounds except 2 and 7. In the presence of dithiothreitol, 4 and 7 formed irreversible complexes with DNA of decreased template properties. The level of the dimers binding was lower than that of the parent compound 2. Cross-links were detected by means of hydroxylapatite chromatography in a complex of the dimer bearing a flexible linking chain, compound 4 with DNA, while the compound 7-DNA complex eluted in the single-stranded DNA region. The extent of cytotoxicity of the two 1-nitro-9-aminoacridine dimers against L1210 cultured cells was different.  相似文献   

20.
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   

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