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 500 毫秒
1.
Measurements of the 1H NMR spectra and relaxation rates were used to study the dynamic properties of 9-aminoacridine (9AA) and four bis(acridine) complexes with d(AT)5.d(AT)5. The behavior of the 9AA (monointercalator) and that of C8 (bisintercalator containing an eight-carbon atom linker chain) are entirely similar. For both compounds, the lifetime of the drug in a particular binding site is 2-3 ms at approximately 20 degrees C, and neither affects the A.T base pair opening rates. The complex with C10 (bisintercalator containing a 10-carbon atom linker chain) is slightly more stable than the C8 complex since its estimated binding site lifetime is 5-10 ms at 29 degrees C. Base pairs adjacent to the bound C10 are destabilized, relative to free d(AT)5.d(AT)5, but other base pairs in the C10 complex are little affected. Bis(acridine) pyrazole (BAPY) and bis(acridine) spermine (BAS) considerably stabilize those base pairs that are sandwiched between the two acridine chromophores, but in the BAS complex proton exchange from the two flanking base pairs appears to be accelerated, relative to free d(AT)5.d(AT)5. The lifetime of these drugs in specific binding sites is too long (>10 ms) to be manifested in increased line widths, at least up to 41 degrees C. An important conclusion from this study is that certain bisintercalators rapidly migrate along DNA, despite having large binding constants (K>10(6) M-1). For C8 and C10 complexes, migration rates are little different from those deduced for 9AA. The rigid linker chain in BAPY and the charge interactions in BAS retard migration of these two bisintercalators. These results provide new parameters that are useful in understanding the biochemical and biological properties of these and other bisintercalating drugs.  相似文献   

2.
We have screened a library of structurally distinct acridine derivatives (19 compounds) for their ability to inhibit lysozyme amyloid aggregation in vitro. Studied acridines were divided into three structurally different groups depending on the molecule planarity and type of the side chain-planar acridines, spiroacridines and tetrahydroacridines. Thioflavine T fluorescence assay and transmission electron microscopy were used for monitoring the inhibiting activity of acridines. We have found that both the structure of the acridine side chains and molecule planarity influence their antiamyloidogenic activity. The planar acridines inhibited lysozyme aggregation effectively. Spiroacridines and tetrahydroacridines had no significant effect on the prevention of lysozyme fibrillization, probably resulting from the presence of the heterocyclic 5-membered ring and non-planarity of molecule. Moreover, in the presence of some tetrahydroacridines the enhanced extent of aggregation was detected. We identified the most active acridine derivates from studied compound library characterized by low micromolar IC(50) values, which indicate their possible application for therapeutic purpose.  相似文献   

3.
Synthetic DNAs were prepared containing 6-methyl adenine (m6A) in place of adenine and 5-ethyl uracil (Et5U) or 5-methoxymethyl uracil (Mm5U) in place of thymine. All three modifications destabilized duplex DNAs to varying degrees. The binding of ethidium was studied to analogues of poly[d(AT)]. There was no evidence of cooperative binding and the "neighbour exclusion rule" was obeyed in all cases although the binding constant to poly[d(m6AT)] was approximately 6 fold higher than to poly[d(AT)]. 31P NMR spectra were recorded in increasing concentrations of CsF. Poly[d(AEt5U)] showed two well-resolved signals separated by 0.55 ppm in 1 M CsF compared to 0.32 ppm for poly[d(AT)] under identical conditions. In contrast, poly[d(AMm5U)] and poly[d(m6AT)] showed two signals separated by 0.28 ppm and 0.15 ppm respectively, only when the concentration of CsF was raised to 2 M. The signals for poly[d(AT)] in 2 M CsF were better resolved and were separated by 0.41 ppm. These results suggest that minor modifications to the bases may have conformational effects which could be recognized by DNA-binding proteins.  相似文献   

4.
Luedtke NW  Liu Q  Tor Y 《Biochemistry》2003,42(39):11391-11403
Semisynthetic aminoglycoside derivatives may provide a means to selectively target viral RNA sites, including the HIV-1 Rev response element (RRE). The design, synthesis, and evaluation of derivatives based upon neomycin B, kanamycin A, and tobramycin conjugates of 9-aminoacridine are presented. To evaluate the importance of the acridine moiety, a series of dimeric aminoglycosides as well as unmodified "monomeric" aminoglycosides have also been evaluated for their nucleic acid affinity and specificity. Fluorescence-based binding assays that use ethidium bromide or Rev peptide displacement are used to quantify the affinities of these compounds to various nucleic acids, including the RRE, tRNA, and duplex DNA. All the modified aminoglycosides exhibit a high affinity for the Rev binding site on the RRE (K(d) 相似文献   

5.
G. Lber  G. Achtert 《Biopolymers》1969,8(5):595-608
The binding constants for the complex formation of more than twenty ring nitrogen-and amino-substituted acridine derivatives with calf thymas DNA were measured by a fluorescence method. DNA quenches the fluorescence of the aminoacridine dyes so long as both amino hydrogens are not substituted. These dyes show an enhancement of their fluorescence intensity in the presence of DNA. Typical representatives of both are proflavine and acridine orange derivatives, respectively. A discussion of steric and electronic influences of various substituents attached to the ring nitrogen and amino groups on the binding led to the concept of different conformations for intercalated acridines without amino groups and the aminoacridines. The electrostatic binding site of the former seems to be the positively charged ring nitrogen, while the binding sites in the aminoacridines are so located that the amino groups are directed towards the negatively charged DNA phosphates.  相似文献   

6.
Acridine-4-carboxamides form a class of known DNA mono-intercalating agents that exhibit cytotoxic activity against tumour cell lines due to their ability to inhibit topoisomerases. Previous studies of bis-acridine derivatives have yielded equivocal results regarding the minimum length of linker necessary between the two acridine chromophores to allow bis-intercalation of duplex DNA. We report here the 1.7 A resolution X-ray crystal structure of a six-carbon-linked bis(acridine-4-carboxamide) ligand bound to d(CGTACG)2 molecules by non-covalent duplex cross-linking. The asymmetric unit consists of one DNA duplex containing an intercalated acridine-4-carboxamide chromophore at each of the two CG steps. The other half of each ligand is bound to another DNA molecule in a symmetry-related manner, with the alkyl linker threading through the minor grooves. The two crystallographically independent ligand molecules adopt distinct side chain interactions, forming hydrogen bonds to either O6 or N7 on the major groove face of guanine, in contrast to the semi-disordered state of mono-intercalators bound to the same DNA molecule. The complex described here provides the first structural evidence for the non-covalent cross-linking of DNA by a small molecule ligand and suggests a possible explanation for the inconsistent behaviour of six-carbon linked bis-acridines in previous assays of DNA bis-intercalation.  相似文献   

7.
We report the first calorimetric investigation of netropsin binding to poly d(AT). Temperature-dependent uv absorption, circular dichroism (CD), batch calorimetry, and differential scanning calorimetry (DSC) were used to detect, monitor, and thermodynamically characterize the binding process. The following results have been obtained: 1) Netropsin groove binding is accompanied by a large exothermic enthalpy of 9.2 kcal/mol of drug bound at 25 degrees C. This indicates that a large negative binding enthalpy may be a necessary but not a sufficient criterion for drug intercalation. We suggest that the exothermic binding might be correlated with specific H-bonding interactions. 2) From the difference in DSC transition enthalpies in the presence and absence of netropsin, we calculate a binding enthalpy of -10.7 kcal/mol of netropsin at 88 degrees C. 3) We calculate a positive delta S for netropsin binding to poly d(AT) at 25 degrees C. This positive entropy change may reflect netropsin-induced release of condensed cations and/or bound water. 4) The netropsin-saturated duplex monophasically melts 46 degrees C higher than the free duplex. The unsaturated duplex melts through two thermally-resolved transitions that correspond to netropsin-free and netropsin-bound regions. These two regions interact dynamically with no substantial influence on the thermal stabilities of the separate domains. 5) Netropsin binding decreases the cooperativity of the duplex to single strand transition.  相似文献   

8.
Symmetrical dimers of lipophilic intercalating chromophores linked by cation-containing chains have recently been shown to have broad-spectrum in vivo anticancer activity. We report the preparation and evaluation of a series of both symmetric and unsymmetric dimers of a variety of intercalating chromophores of varied DNA binding strength, including naphthalimides, acridines, phenazines, oxanthrenes and 2-phenylquinolines. The unsymmetrical dimers were prepared by sequential coupling of the chromophores to linkers with selectively protected primary terminal amines to ensure high yields and unequivocal product. Protection of the internal (secondary) amines as BOC derivatives was used to ensure complete structural specificity, and was also an aid to the purification of these very polar compounds. The growth inhibitory abilities (as IC(50) values) of the compounds in a range of cell lines showed that the nature of the linker chain was important, and independent of the nature of the chromophore, with compounds containing the dicationic linker [-(CH2)2NH(CH2)2NH(CH2)2-] being on average 30-fold more potent than the corresponding compounds containing the monocationic linker [-(CH2)3NMe(CH2)3-]. However, the chromophores also play a role in determining biological activity, with the cytotoxicities of symmetric and unsymmetric dicationic dimers correlating with the overall DNA binding abilities of the chromophores.  相似文献   

9.
Structure of a pleiomeric form of poly d(AT):poly d(AT)   总被引:1,自引:2,他引:1       下载免费PDF全文
A chemically simple polynucleotide duplex, poly d(AT):poly d(AT), has been trapped in a fibrous form with a complex helical secondary structure with a large (7.4 nm) axial repeat 24 nucleotides long. The motif which is repeated by the symmetry elements is a hexanucleotide in which two residues (both TpA) have the less common gauche minus conformation at C3'-O3' and consequently distinctive phosphate orientations. This reinforces earlier conclusions that PypPu nucleotides tend to have different shapes from PupPy nucleotides and that DNA surfaces may signal what base sequences lie beneath them. The morphological differences between this pleiomeric DNA polymer and closely-related, but more symmetrical allomorphs are just as great as those observed in short DNA fragments in crystals.  相似文献   

10.
The photophysics of the sequence isomers d(TA) and d(AT) has been investigated at room temperature in 5 x 10(-5) M neutral aqueous solution using pulsed ultraviolet excitation from the ACO synchrotron and detection by time correlation or gated single-photon counting. Decay profiles of the emissions at 350, 400 and 460 have been analyzed both independently and globally by reiterative non-linear least-squares fitting to models of two and three independently emitting species. No evidence has been observed for excited-state reaction. Time-windowed spectra, both emission and excitation, have been collected for three time windows and have been deconvoluted to give time-resolved spectra using the lifetimes resulting from the decay analyses. Spectra are separated into two classes, with picosecond and nanosecond lifetimes, respectively. The picosecond spectra have the emission and excitation spectral characteristics of mixed monomer (A and T) fluorescences and are assigned as originating from the unstacked fractions of d(TA) and d(AT). The nanosecond emission spectra from d(TA) and d(AT) are both two-component, with lambda max approximately 350 and approximately 425 nm and lifetimes of 2.3 and 6.1 ns, respectively. The time-resolved excitation spectra for the nanosecond emissions are quite different from the isotropic absorption spectra of d(TA) and d(AT) but correlate with the anisotropic absorption for out-of-plane transitions between stacked bases of co-crystals of 9-methyladenine and 1-methylthymine reported by Stewart and Davidson. The nanosecond spectra thus represent the direct excitation and emission of stacked pairs of bases. These results provide no evidence for energy transfer and are probably related to sequence-specific photo-adduct formation.  相似文献   

11.
A structural transition in d(AT)n.d(AT)n inserts within superhelical DNA   总被引:7,自引:0,他引:7  
We have constructed plasmids carrying d(AT)n.d(AT)n inserts of different lengths. Two-dimensional gel electrophoresis patterns show that an increase in the negative superhelicity of these DNAs brings about a structural transition within the inserts, resulting in a reduction of the superhelical stress. However, this reduction corresponds to the expected values neither for cruciform nor the Z form. Those DNA topoisomers in which the structural transition had occurred proved to be specifically recognizable by single-strand-specific endonuclease S1, with the cleavage site situated at the centre of the insert. These data, as well as kinetic studies, suggest that the cloned d(AT)n.d(AT)n sequences adopt a cruciform rather than the Z-form structure. We discuss plausible reasons of the discrepancy between the observed superhelical stress release and that expected for the transition of the insert to the cruciform state.  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.
We report the first calorimetric investigation of steroid diamine binding to a DNA duplex. Absorption spectroscopy, batch calorimetry, and differential scanning calorimetry (DSC) have been used to detect, monitor, and thermodynamically characterize the binding of the steroid diamine, dipyrandium, to poly d(AT). The following thermodynamic data for the binding in 16 mM Na+ at 25 degrees C have been obtained: delta G degree = -6.5 kcal/mol, delta H degree = +4.2 kcal/mol, and delta S = +36 e.u. We interpret the endothermic binding enthalpy in terms of steroid-induced conformational changes in the duplex (e.g. "kinking"). The large positive entropy is interpreted in terms of binding-induced release of bound water and condensed sodium ions. The salt-dependence of the binding constant is interpreted in terms of dipyrandium site-binding involving only one of the two charged ends of the steroid. The optical and DSC curves for the unsaturated steroid-poly d(AT) complexes exhibit biphasic behavior. A comparison of the van't Hoff and the calorimetric transition enthalpies reveals that steroid binding reduces the cooperativity of the transition.  相似文献   

14.
A trimer made up of three acridine chromophores linked by a positively charged poly(aminoalkyl) chain was synthesized as a potential tris-intercalating agent. The length of the linking chain was selected to allow intercalation of each chromophore according to the excluded site model. 1H NMR studies have shown that, at 5 mM sodium, pH 5, the acridine trimer occurred under a folded conformation stabilized by stacking interactions between the three aromatic rings. DNA tris-intercalation of the dye at a low dye/base pair ratio was shown by measurements of both the unwinding of PM2 DNA and the lengthening of sonicated rodlike DNA. The trimer exhibits a high DNA affinity for poly[d(A-T)] (Kapp = 8 X 10(8) M-1, 1 M sodium) as shown by competition experiments with ethidium dimer. Kinetic studies of both the association with poly[d(A-T)] and the exchange between poly[d(A-T)] and sonicated calf thymus DNA have been performed as a function of the ionic strength. In 0.3 M sodium the on-rate constant (k1 = 2.6 X 10(7) M-1 s-1) is similar to that reported for other monoacridines or bis(acridines), whereas the off-rate constant is much smaller (k-1 = 1.2 X 10(-4) s-1), leading to an equilibrium binding constant as large as Kapp = 2.2 X 10(11) M-1. A plot of log (k1/k-1) as a function of log [Na+] yielded a straight line whose slope shows that 5.7 ion pairs (out of 7 potential) are formed upon the interaction with DNA. From this linear relationship a Kapp value of 10(14) M-1 in 0.1 M sodium can be estimated. Such a value reaches and even goes beyond that of some DNA regulatory proteins. This acridine trimer appears to be the first synthetic ligand with such a high DNA affinity.  相似文献   

15.
The molecular basis of the marked structure-activity relationship for a homologous series of DNA-binding phenoxazone drugs (ActII-ActIV) has been investigated by NMR spectroscopy and molecular mechanics. The spatial structures of the complexes between the drugs and a model deoxytetranucleotide, 5'-d(TpGpCpA), have been determined by molecular mechanics methods using homonuclear (1)H-(1)H 2D-NOESY and heteronuclear (1)H-(31)P (HMBC) NMR spectroscopic data. Observed intermolecular NOE contacts and equilibrium binding studies confirm that the binding affinity of the synthetic phenoxazone derivatives with d(TGCA) decreases with an increase in the number of CH(2) groups in the dimethylaminoalkyl side chains, i.e., ActII > ActIII > ActIV, in agreement with the observed biological activity of these compounds. Molecular mechanics calculations of the spatial structures of the intercalated complexes of ActII-ActIV with d(TGCA) indicate that the different binding constants of the phenoxazone derivatives with the DNA oligomer are due to the different degrees of intercalation of the chromophore and the different steric arrangements of aminoalkyl side chains in the minor groove of the tetramer duplex; this results in different distances between the negatively-charged phosphates of the DNA duplex and the terminal positively-charged N(CH(3))(2) groups of the side chains.  相似文献   

16.
Coralyne and several other synthetic benzo[a,g]quinolizium derivatives related to protoberberine alkaloids have exhibited activity as topoisomerase poisons. These compounds are characterized by the presence of a positively charged iminium group, which has been postulated to be associated with their pharmacological properties. The objective of the present study was to devise stable noncharged bioisosteres of these compounds. Several similarly substituted benz[a]acridine and benz[c]acridine derivatives were synthesized and their relative activity as topoisomerase poisons was determined. While the benz[c]acridine derivatives evaluated as part of this study were devoid of topoisomerase poisoning activity, several dihydrobenz[a]acridines were able to enhance DNA cleavage in the presence of topo I. In contrast to certain protoberberine derivatives that did exhibit activity as topo II poisons, none of the benz[a]acridines derivatives enhanced DNA cleavage in the presence of topo II. Among the benz[a]acridines studied, 5,6-dihydro-3,4-methylenedioxy-9,10-dimethoxybenz[a]acridine, 13e, was the most potent topo I poison, with comparable potency to coralyne. These data suggest that heterocyclic compounds structurally related to coralyne can exhibit potent topo I poisoning activity despite the absence of an iminium cation within their structure. In comparison to coralyne or other protoberberine derivatives, these benz[a]acridine derivatives possess distinctly different physicochemical properties and represent a novel series of topo I poisons.  相似文献   

17.
Previous studies have demonstrated that the internalization of the angiotensin II type 1A receptor (AT(1A)R) may be mediated by both beta-arrestin-sensitive and -insensitive mechanisms. Therefore, we have used the AT(1A)R carboxyl-terminal tail to screen a rat brain yeast two-hybrid expression library for novel AT(1A)R-interacting proteins that might contribute to the regulation of AT(1A)R internalization. We have identified Rab5a as an AT(1A)R-binding protein that selectively associates with the AT(1A)R and not with the beta2-adrenergic receptor. A Rab5a-S34N mutant defective in GTP binding does not prevent the internalization of the AT(1A)R but does prevent the trafficking of the AT(1A)R into larger hollow cored vesicular structures. Agonist activation of the AT(1A)R promotes both the formation of Rab5a.AT(1A)R protein complexes and Rab5a GTP binding. Rab5a interactions with the AT(1A)R are mediated in part by the last 10 amino acid residues of the AT(1A)R carboxyl-terminal tail, and although a mutant receptor lacking these residues internalizes normally, it does not redistribute into larger hollow vesicles. Our data suggest that AT(1A)R activation modulates Rab5a activity leading to the homotypic fusion of endocytic vesicles. These observations suggest that vesicular cargo proteins, such as the AT(1A)R, may control their targeting between intracellular compartments by directly regulating the activity of components of the intracellular trafficking machinery such as Rab5a.  相似文献   

18.
Bridged bis(beta-cyclodextrin) 1 with a pyridine-2,6-dicarboxamide linker was synthesized, and its inclusion complexation behavior with some aliphatic oligopeptides was investigated in aqueous buffer solution of pH 2.0 and 7.2 at 25 degrees C by means of circular dichroism, fluorescence, and 2D NMR techniques. The results show that the resulting inclusion complexes of 1 with oligopeptides adopt a cooperative "cyclodextrin-guest-cyclodextrin" sandwich binding mode in a neutral media, but a "guest-linker-cyclodextrin" coinclusion binding mode in an acidic media. These switchable binding modes consequently rationalize the binding ability of bis(beta-cyclodextrin) 1 at different pH values; that is, 1 shows the stronger association with oligopeptides in a neutral media. Because of the simultaneous contributions of hydrophobic, hydrogen bond, and electrostatic interactions, bis(beta-cyclodextrin) 1 affords length-selectivity up to 4.7 for the Gly-Gly/Gly-Gly-Gly pair at pH 2.0 and sequence-selectivity up to 4.2 for the Gly-Leu/Leu-Gly pair at pH 7.2. These phenomena are discussed from the viewpoint of the size-fit concept and the multipoint recognitions between host and guest.  相似文献   

19.
The macrocyclic bisacridine (CBA) has been reported previously to specifically recognize single-stranded nucleic acid structures, especially DNA hairpins. The binding of the drug with an abasic site-containing oligonucleotide, was investigated by (1)H NMR and molecular modeling. We have used a DNA undecamer, the d(C(1)G(2)C(3)A(4)C(5)X(6)C(7)A(8)C(9)G(10)C(11)) x d(G(12)C(13)G(14)T(15)G(16)T(17)G(18)T(19)G(2)(0)C(21)G(22)) duplex in which the X residue is a stable analogue of the abasic site [3-hydroxy-2-(hydroxymethyl) tetrahydrofuran]. Analysis of the NMR data reveals that the bisacridine molecule forms two different intercalation complexes in a 80/20 (+/- 10) ratio. For the major complex, a molecular modeling study was performed guided by nineteen intermolecular drug-DNA restraints, determined from NOESY spectra. In this model, the ligand interacts in the threading binding mode with an acridine ring intercalated between the C(7)-A(8) and T(15)-G(16) base pairs, while the other acridine ring resides in the abasic pocket. The two linker chains are positioned in the minor and in the major groove, respectively. A comparable study was performed to evaluate the interaction of CBA with the parent unmodified duplex in which X(6) was replaced by an adenine residue. No complex formation was observed when operating in identical conditions. This shows the selective binding of CBA to the abasic site and its potential interest to target the abasic site lesion.  相似文献   

20.
Chemical cross-linking was used to study the interaction between non-histone high-mobility-group (HMG)1 and histone H5 in free solution. The presence of acidic C-terminal domain in HMG1 was shown to be a prerequisite for HMG1 binding to histone H5. The objective of this communication is to ascertain whether HMG1 could affect the conformation of DNA associated with a linker histone H5. Complexes of histone H5 with chicken erythrocyte DNA or an alternating purine-pyrimidine polynucleotide poly[d(A-T)] were prepared at different molar ratios H5/DNA. Changes in DNA conformation in the complexes with histone H5 or H5/HMG1 were monitored by circular dichroism (c.d.). Depending on the molar ratio H5/poly[d(A-T)], under conditions limiting the complex aggregation, three distinct types of c.d. spectra were observed. The addition of HMG1 to H5-DNA complexes reduced in all cases the histone H5-induced conformational changes in poly[d(A-T)]. The sensitivity of H5-poly[d(A-T)] complexes to HMG1 was inversely proportional to the amount of H5 in the complex. The effect of HMG1 was not observed upon removal of the acidic C-terminal domain of HMG1.  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