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 31 毫秒
1.
Base-stacking is a key factor in the energetics that determines nucleic acid structure. We measure the tensile response of single-stranded DNA as a function of sequence and monovalent salt concentration to examine the effects of base-stacking on the mechanical and thermodynamic properties of single-stranded DNA. By comparing the elastic response of highly stacked poly(dA) and that of a polypyrimidine sequence with minimal stacking, we find that base-stacking in poly(dA) significantly enhances the polymer’s rigidity. The unstacking transition of poly(dA) at high force reveals that the intrinsic electrostatic tension on the molecule varies significantly more weakly on salt concentration than mean-field predictions. Further, we provide a model-independent estimate of the free energy difference between stacked poly(dA) and unstacked polypyrimidine, finding it to be ∼−0.25 kBT/base and nearly constant over three orders of magnitude in salt concentration.  相似文献   

2.
The photocleavage of double-stranded and single-stranded DNA by the fluorescent dye YOYO-1 was investigated in real time by using the synchrotron radiation light source ASTRID (ISA, Denmark) both to initiate the reaction and to monitor its progress using Couette flow linear dichroism (LD) throughout the irradiation period. The dependence of LD signals on DNA sequences and on time in the intense light beam was explored and quantified for single-stranded poly(dA), poly[(dA-dT)2], calf thymus DNA (ctDNA) and Micrococcus luteus DNA (mlDNA). The DNA and ligand regions of the spectrum showed different LD kinetic behaviors, and there was significant sequence dependence of the kinetics. However, in contrast to expectations from the literature, we found that poly(dA), mlDNA, low salt ctDNA and low salt poly[(dA-dT)2] all had significant populations of groove-bound YOYO. It seems that this mode was predominantly responsible for the catalysis of DNA cleavage. In homopolymeric DNAs, intercalated YOYO was unable to cleave DNA. In mixed-sequence DNAs the data suggest that YOYO in some but not all intercalated binding sites can cause cleavage. It is also likely that cleavage occurs at transient single-stranded regions. The reaction rates for a 100 mA beam current of 0.5-μW power varied from 0.6 h−1 for single-stranded poly(dA) to essentially zero for low salt poly[(dG-dC)2] and high salt poly[(dA-dT)2]. At the conclusion of the experiments with each kind of DNA, uncleaved DNA with intercalated YOYO remained.  相似文献   

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

4.
The effect that Escherichia coli single-stranded DNA binding (SSB) protein has on the single-stranded DNA-dependent ATPase activity of RecA protein is shown to depend upon a number of variables such as order of addition, magnesium concentration, temperature and the type of single-stranded DNA substrate used. When SSB protein is added to the DNA solution prior to the addition of RecA protein, a significant inhibition of ATPase activity is observed. Also, when SSB protein is added after the formation of a RecA protein-single-stranded DNA complex using either etheno M13 DNA, poly(dA) or poly(dT), or using single-stranded phage M13 DNA at lower temperature (25 °C) and magnesium chloride concentrations of 1 mm or 4 mm, a time-dependent inhibition of activity is observed. These results are consistent with the conclusion that SSB protein displaces the RecA protein from these DNA substrates, as described in the accompanying paper. However, if SSB protein is added last to complexes of RecA protein and single-stranded M13 DNA at elevated temperature (37 °C) and magnesium chloride concentrations of 4 mm or 10 mm, or to poly(dA) and poly(dT) that was renatured in the presence of RecA protein, no inhibition of ATPase activity is observed; in fact, a marked stimulation is observed for single-stranded M13 DNA. A similar effect is observed if the bacteriophage T4-coded gene 32 protein is substituted for SSB protein. The apparent stoichiometry of DNA (nucleotides) to RecA protein at the optimal ATPase activity for etheno M13 DNA, poly(dA) and poly(dT) is 6(±1) nucleotides per RecA protein monomer at 4 mm-MgCl2 and 37 °C. Under the same conditions, the apparent stoichiometry obtained using single-stranded M13 DNA is 12 nucleotides per RecA protein monomer; however, the stoichiometry changes to 4.5 nucleotides per RecA protein monomer when SSB protein is added last. In addition, a stoichiometry of four nucleotides per RecA protein can be obtained with single-stranded M13 DNA in the absence of SSB protein if the reactions are carried out in 1 mm-MgCl2. These data are consistent with the interpretation that secondary structure within the natural DNA substrate limits the accessibility of RecA protein to these regions. The role of SSB protein is to eliminate this secondary structure and allow RecA protein to bind to these previously inaccessible regions of the DNA. In addition, our results have disclosed an additional property of the RecA protein-single-stranded DNA complex: namely, in the presence of complementary base-pairing and at elevated temperatures and magnesium concentrations, a unique RecA protein-DNA complex forms that is resistant to inhibition by SSB protein.  相似文献   

5.
Melting measurements of the mono-base-pair DNA polymers showed that the melting temperature Tm of the B-DNA homopolymer poly (dA ) · poly (dT) is higher than that of the copolymer poly [d(A-T)]. On the other hand, the Tmof the B-DNA homopolymer poly (dG) · poly (dC) is lower than that of the copolymer poly [d (G-C)]. From a structural point of view, the cross-strand base-stacking interaction in a DNA homopolymer is weaker than that in a DNA copolymer with the same base pair. One would then expect that all the DNA homopolymers are less stable than the copolymer with the same base pair. We find that the inversion of the melting order seen in the AT mono-base-pair DNA polymers is caused by the enhanced thermal stability of poly (dA) · poly (dT) from a well-defined spine of hydration attached to its minor groove. In this paper we employ the modified self-consistent phonon theory to calculate base-pair opening probabilities of four B-DNA polymers: poly(dA)-poly(dT), poly(dG) · poly(dC), poly[d(A-T)], and poly[d(G-C)] at temperatures from room temperature through the melting regions. Our calculations show that the spine of hydration can give the inverted melting order of the AT polymers as compared to the GC polymers in fair agreement with experimental measurements. Our calculated hydration spine disruption behavior in poly(dA) · poly(dT) at premelting temperatures is also in agreement with experimentally observed premelting transitions in poly (dA) · poly (dT). The work is in a sense a test of the validity of our models of nonbonded interactions and spine of hydration interactions. We find we have to develop the concept of a strained bond to fit observations in poly (dA) · poly(dT). The strained-bond concept also explains the otherwise anomalous stability of the hydration chain. © 1993 John Wiley & Sons, Inc.  相似文献   

6.
The equilibrium binding of the cytotoxic plant alkaloid berberine to various DNAs and energetics of the interaction have been studied. At low ratios of bound alkaloid to base pair, the binding exhibited cooperativity to natural DNAs having almost equal proportions of AT and GC sequences. In contrast, the binding was non-cooperative to DNAs with predominantly high AT or GC sequences. Among the synthetic DNAs, cooperative binding was observed with poly(dA).poly(dT) and poly(dG).poly(dC) while non-cooperative binding was seen with poly(dA–dT).poly(dA–dT) and poly(dG–dC).poly(dG–dC). Both cooperative and non-cooperative bindings were remarkably dependent on the salt concentration of the media. Linear plots of ln Ka versus [Na+] for poly(dA).poly(dT) and poly(dA–dT).poly(dA–dT) showed the release of 0.56 and 0.75 sodium ions respectively per bound alkaloid. Isothermal titration calorimetry results revealed the binding to be exothermic and favoured by both enthalpy and entropy changes in all DNAs except the two AT polymers and AT rich DNA, where the same was predominantly entropy driven. Heat capacity values (ΔCpo) of berberine binding to poly(dA).poly(dT), poly(dA–dT).poly(dA–dT), Clostridium perfringens and calf thymus DNA were − 98, − 140, − 120 and − 110 cal/mol K respectively. This study presents new insights into the binding dependent base pair heterogeneity in DNA conformation and the first complete thermodynamic profile of berberine binding to DNAs.  相似文献   

7.
Mou TC  Gray CW  Gray DM 《Biophysical journal》1999,76(3):1537-1551
The Ff gene 5 protein (g5p) is considered to be a nonspecific single-stranded DNA binding protein, because it binds cooperatively to and saturates the Ff bacteriophage single-stranded DNA genome and other single-stranded polynucleotides. However, the binding affinity Komega (the intrinsic binding constant times a cooperativity factor) differs by over an order of magnitude for binding to single-stranded polynucleotides such as poly[d(A)] and poly[d(C)]. A polynucleotide that is more stacked, like poly[d(A)], binds more weakly than one that is less stacked, like poly[d(C)]. To test the hypothesis that DNA base stacking, a nearest-neighbor property, is involved in the binding affinity of the Ff g5p for different DNA sequences, Komega values were determined as a function of NaCl concentration for binding to six synthetic sequences 48 nucleotides in length: dA48, dC48, d(AAC)16, d(ACC)16, d(AACC)12, and d(AAACC)9A3. The binding affinities of the protein for these sequences were indeed found to be related to the nearest-neighbor compositions of the sequences, rather than to simple base compositions. That is, the g5p binding site, which is spanned by four nucleotides, discriminates among these sequences on the basis of the relative numbers of nearest neighbors (AA, CC, and AC plus CA) in the sequence. The results support the hypothesis that the extent of base stacking/unstacking of the free, nonbound ssDNA plays an important role in the binding affinity of the Ff gene 5 protein.  相似文献   

8.
M-DNA is a complex between the divalent metal ions Zn2+, Ni2+ and Co2+ and duplex DNA which forms at a pH of ~8.5. The stability and formation of M-DNA was monitored with an ethidium fluorescence assay in order to assess the relationship between pH, metal ion concentration, DNA concentration and the base composition. The dismutation of calf thymus DNA exhibits hysteresis with the formation of M-DNA occurring at a higher pH than the reconversion of M-DNA back to B-DNA. Hysteresis is most prominent with the Ni form of M-DNA where complete reconversion to B-DNA takes several hours even in the presence of EDTA. Increasing the DNA concentration leads to an increase in the metal ion concentration required for M-DNA formation. Both poly(dG)•poly(dC) and poly(dA)•poly(dT) formed M-DNA more readily than the corresponding mixed sequence DNAs. For poly(dG)•(poly(dC) M-DNA formation was observed at pH 7.4 with 0.5 mM ZnCl2. Modified bases were incorporated into a 500 bp fragment of phage λ DNA by polymerase chain reaction. DNAs in which guanine was replaced with hypoxanthine or thymine with 5-fluorouracil formed M-DNA at pHs below 8 whereas substitutions such as 2-aminoadenine and 5-methylcytosine had little effect. Poly[d(A5FU)] also formed a very stable M-DNA duplex as judged from Tm measurements. It is evident that the lower the pKa of the imino proton of the base, the lower the pH at which M-DNA will form; a finding that is consistent with the replacement of the imino proton with the metal ion.  相似文献   

9.
The interaction of the fluorescent dye thiazole orange (TO) with nucleic acids is characterized. It is found that TO binds with highest affinity to double-stranded (ds) DNA [log(K) ≈ 5.5 at 100 mM salt], about 5–10 times weaker to single-stranded polypurines, and further 10–1000 times weaker to single-stranded polypyrimidines. TO binds as a monomer to dsDNAs and poly(dA), both as a monomer and as a dimer to poly(dG) and mainly as a dimer to poly(dC) and poly(dT). The fluorescence quantum yield of TO free in solution is about 2 · 10−4, and it increases to about 0.1 when bound to dsDNA or to poly(dA), and to about 0.4 when bound to poly(dG). Estimated quantum yields of TO bound to poly(dC) and poly(dT) are about 0.06 and 0.01, respectively. The quantum yield of bound TO depends on temperature and decreases about threefold between 5 and 50°C. © 1998 John Wiley & Sons, Inc. Biopoly 46: 39–51, 1998  相似文献   

10.
In this paper we examine the specificity of the co-operative binding (in the polynucleotide mode) of bacteriophage T4-coded gene 32 protein to synthetic and natural single-stranded nucleic acids differing in base composition and sugar type. It is shown by competition experiments in a tight-binding (low salt) environment that there is a high degree of binding specificity under these (protein-limiting) conditions, with one type of nucleic acid lattice binding gene 32 protein to saturation before any binding to the competing lattice takes place; it is also shown that the same differential specificities apply at high salt concentrations. Procedures developed in the preceding paper (Kowalczykowski et al., 1980) are used to measure the net binding affinities () of gene 32 protein to a variety of polynucleotides, as well as to determine individual values of K and ω for some systems. For all polynucleotides, virtually the entire specificity and salt dependence of binding of appears to be in K. In ~0.2 m-NaCl, the net binding affinities () range from ~106 to ~1011m?1; in order of increasing affinities we find: poly(rC) < poly(rU) < poly(rA) < poly(dA) < poly(dC) < poly(dU) < poly(rI) < poly(dI) < poly-(dT). In general, for a particular homopolyribonucleotide at constant salt concentration is 101 to 104smaller than for the corresponding homopoly-deoxyribopolynucleotide. Values of for randomly copolymerized polynucleotides and for natural DNA fall at the compositionally weighted average of the values for the individual homopolynucleotides (except for poly(dT), which appears to bind somewhat tighter), indicating that the net affinity represents the sum of the binding free energy contributions of the individual nucleotides. It is shown that these results, on a competition basis under physiological salt conditions, can account quantitatively for the autogenous regulation of the synthesis of gene 32 protein at the translational level (Russel et al., 1976; Lemaire et al., 1978). In addition, these results suggest possible mechanisms by which gene 32 messenger RNA might be specifically recognized (by gene 32 protein) and functionally discriminated from the other mRNAs of phage T4.  相似文献   

11.
Origin of the intrinsic rigidity of DNA   总被引:3,自引:2,他引:1       下载免费PDF全文
The intrinsic rigidities of DNA and RNA helices are generally thought to arise from some combination of vertical base-stacking interactions and intra-helix phosphate–phosphate charge repulsion; however, the relative contributions of these two types of interaction to helix rigidity have not been quantified. To address this issue, we have measured the rotational decay times of a ‘gapped-duplex’ DNA molecule possessing a central, single-stranded region, dT24, before and after addition of the free purine base, N6-methyladenine (meA). Upon addition of meA, the bases pair with the T residues, forming a continuous stack within the gap region. Formation of the gapped duplex is accompanied by a nearly 2-fold increase in decay time, to values that are indistinguishable from the full duplex control for monovalent salt concentrations up to 90 mM. These results indicate that at least 90% of the rigidity of the dTn–dAn homopolymer derives from base pair stacking effects, with phosphate–phosphate interactions contributing relatively little to net helix rigidity at moderate salt concentrations.  相似文献   

12.
Native DNA duplexes in fibers exist usually in one of three well-known (A, B and C) forms depending on relative humidity, type of cations and the amount of retained salt. To determine the precise influence of these factors and the effect of base composition, as well as base sequence, on DNA secondary structure, X-ray diffraction methods have been used to study all four synthetic DNA duplexes with repeated dinucleotide sequences, eight of the 12 with repeated trinucleotide sequences and seven analogues in which guanine was replaced with hypoxanthine. The results indicate that there are at least six additional allomorphs denoted by B′, C′, C″, D, E and S.The B′ form (h = 0.329 nm) observed for poly(dA) · poly(dT), poly(dI) · poly(dC) and poly[d(A-I)] · poly[d(C-T)] is a minor variant of the traditional B form (h = 0.338 nm) of native DNA. The two C-like forms C′ for poly[d(A-G-C)] · poly-[d(G-C-T)] and poly[d(G-G-T)] · poly[d(A-C-C)] and C″ for poly[d(A-G)] · poly-[d(C-T)] have, respectively, 91 and 92 symmetries which reflect repetition of trinucleotide and dinucleotide sequences, respectively. Although isocompositional with poly(dA) · poly(dT), the existence of the rather different D form (81) for poly[d(A-T)] · poly[d(A-T)] or for poly[d(A-A-T)] · poly[d(A-T-T)] is a clear demonstration of the sequence effect. The I · C pair generally mimics an A · T pair, but poly[d(I-I-T)] · poly[d(A-C-C)] provides a new (E) form with approximately 152 screw symmetry and with 〈h〉 = 0.325 nm and 〈t〉 = 48 dg per nucleotide. The S form (65) observed for poly[d(G-C)] · poly[d(G-C)] and poly[d(A-C)] · poly[d(G-T)] is an unusual left-handed polydinucleotide helix and is accessible to any alternating purine-pyrimidine sequence. In it the two nucleotides have quite different conformations and involve syn purine and anti pyrimidine nucleosides.  相似文献   

13.
Jain SS  Polak M  Hud NV 《Nucleic acids research》2003,31(15):4608-4615
Small molecules that intercalate in DNA and RNA are powerful agents for controlling nucleic acid structural transitions. We recently demonstrated that coralyne, a small crescent-shaped molecule, can cause the complete and irreversible disproportionation of duplex poly(dA)·poly(dT) into triplex poly(dA)·poly(dT)·poly(dT) and a poly(dA) self- structure. Both DNA secondary structures that result from duplex disproportionation are stabilized by coralyne intercalation. In the present study, we show that the kinetics and thermodynamics of coralyne-driven duplex disproportionation strongly depend on oligonucleotide length. For example, disproportionation of duplex (dA)16·(dT)16 by coralyne reverts over the course of hours if the sample is maintained at 4°C. Coralyne-disproportioned (dA)32· (dT)32, on the other hand, only partially reverts to the duplex state over the course of days at the same temperature. Furthermore, the equilibrium state of a (dA)16·(dT)16 sample in the presence of coralyne at room temperature contains three different secondary structures [i.e. duplex, triplex and the (dA)16 self-structure]. Even the well-studied process of triplex stabilization by coralyne binding is found to be a length-dependent phenomenon and more complicated than previously appreciated. Together these observations indicate that at least one secondary structure in our nucleic acid system [i.e. duplex, triplex or (dA)n self-structure] binds coralyne in a length-dependent manner.  相似文献   

14.
Diadenosine 5′,5‴-P1,P4-tetraphosphate (Ap4A) stimulates DNA synthesis in Xenopus laevis oocytes in the presence of activated DNA as template. Besides Ap4A, other analogues such as Ap3A, ATP and other derivatives are able to stimulate DNA polymerase activity. The effect of Ap4A on DNA synthesis is observed with poly(dT) and poly(dT)-poly(dA) as templates, while no effect is found with poly(dA)(dT)12–18 and poly(dC)(dG)12–18. In the presence of a poly(dT) template, the oocyte extract is able to utilize Ap4A as primer and to form a covalent bond between this dinucleotide and the nascent poly(dA) chain. An Ap4A-binding protein present in the system has been purified and separated from DNA polymerase α-primase after phosphocellulose chromatography. After this separation, Ap4A is no longer able to stimulate the polymerase activity, or to be utilized as primer by DNA polymerase α-primase.  相似文献   

15.
To identify and assess the specificity of the 1,25-dihydroxyvitamin D3 chick intestinal cytoplasmic receptor's nucleotide binding site, a competitive DNA-cellulose binding assay was utilized. Unlike other steroid hormone receptors, the 1,25-dihydroxyvitamin D3-receptor complex binds homologous DNA at 4 °C and does not appear to undergo thermal- or salt-induced activation. Data are presented which suggest that receptor binding discriminates between double-stranded DNA and RNA but is not specific with respect to DNA base sequences. However, DNA base sequence selectivity by 1,25-dihydroxyvitamin D3-receptor complexes is observed using synthetic polydeoxyribonucleotides, particularly, poly(dA-dT) · poly(dA-dT) and poly(dA) · poly(dT). Preference for double-stranded over single-stranded DNA was also observed. Consistent with this finding, both actinomycin D and ethidium bromide caused a dose-dependent inhibition of receptor binding to DNA-cellulose. It is concluded that the 1,25-dihydroxyvitamin D3-receptor complex has specificity for AT-rich segments of double-stranded DNA and that this interaction is not merely electrostatic, but also involves hydrophobic interaction with the major and/or minor grooves of the DNA helix.  相似文献   

16.
The binding characteristics of the interaction of gene 5 protein with polynucleotides, i.e. poly(dA), poly(dT) and M13 DNA, have been determined by following the quenching of the protein fluorescence. In general, the binding is highly co-operative and for the binding of the protein to poly(dA) and M13 DNA the co-operativity parameter ω is estimated to have values between 50 and 300. Under comparable experimental conditions, the intrinsic binding constant Kint is at least two orders of magnitude higher for poly(dT) than for poly(dA), while the value for M13 DNA is intermediate. For poly(dA), the binding has been studied as a function of ionic strength and temperature. From these experiments it can be concluded that ionic interactions as well as van der Waals interactions (e.g. stacking interactions) are important for the complex formation of the protein with polynucleotides. From a comparison of the binding of the protein to poly(dA) and poly(dT), it is concluded that stacking interactions in the polynucleotide have a negative influence on protein binding. This conclusion, in conjunction with the weak temperature dependence of Kint. indicates that ionic interactions play a major role in the stabilization of the protein-poly(dA) complex. The co-operativity factor ω is little or not dependent on the ionic strength or the type of polynucleotide involved in binding. It is determined by interactions between complexed protein molecules. These interactions are primarily non-electrostatic.The binding characteristics obtained for the gene 5 protein-polynucleotide complexes are compared with those we have found for the binding to small oligonucleotides. It appears that oligonucleotide and polynucleotide binding differ in many aspects; i.e. there is a difference in Kint, ω and the number of nucleotides covered. The validity of linear lattice binding theories is discussed in this context. By comparing the binding parameters found for the gene 5 protein with those of the Escherichia coli DNA binding protein I. it is possible to explain the displacement of the E. coli protein by the gene 5 protein that occurs in vivo.  相似文献   

17.
Abstract

Thermodynamic parameters of melting process (δHm, Tm, δTm) of calf thymus DNA, poly(dA)poly(dT) and poly(d(A-C))·poly(d(G-T)) were determined in the presence of various concentrations of TOEPyP(4) and its Zn complex. The investigated porphyrins caused serious stabilization of calf thymus DNA and poly poly(dA)poly(dT), but not poly(d(A-C))poly(d(G-T)). It was shown that TOEpyp(4) revealed GC specificity, it increased Tm of satellite fraction by 24°C, but ZnTOEpyp(4), on the contrary, predominately bound with AT-rich sites and increased DNA main stage Tm by 18°C, and Tm of poly(dA)poly(dT) increased by 40 °C, in comparison with the same polymers without porphyrin. ZnTOEpyp(4) binds with DNA and poly(dA)poly(dT) in two modes—strong and weak ones. In the range of r from 0.005 to 0.08 both modes were fulfilled, and in the range of r from 0.165 to 0.25 only one mode—strong binding—took place. The weak binding is characterized with shifting of Tm by some grades, and for the strong binding Tm shifts by ~ 30–40°C. Invariability of ΔHm of DNA and poly(dA)poly(dT), and sharp increase of Tm in the range of r from 0.08 to 0.25 for thymus DNA and 0.01–0.2 for poly(dA)poly(dT) we interpret as entropic character of these complexes melting. It was suggested that this entropic character of melting is connected with forcing out of H2O molecules from AT sites by ZnTOEpyp(4) and with formation of outside stacking at the sites of binding. Four-fold decrease of calf thymus DNA melting range width ΔTm caused by increase of added ZnTO- Epyp(4) concentration is explained by rapprochement of AT and GC pairs thermal stability, and it is in agreement with a well-known dependence, according to which ΔT~TGC-TAT for DNA obtained from higher organisms (L. V. Berestetskaya, M. D. Frank-Kamenetskii, and Yu. S. Lazurkin. Biopolymers 13, 193–205 (1974)). Poly (d(A-C))poly(d(G-T)) in the presence of ZnTOEpyp(4) gives only one mode of weak binding. The conclusion is that binding of ZnTOEpyp(4) with DNA depends on its nucleotide sequence.  相似文献   

18.
The binding site and the geometry of Co(III)meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (CoTMPyP) complexed with double helical poly(dA)·poly(dT) and poly(dG)·poly(dC), and with triple helical poly(dA)·[poly(dT)]2 and poly(dC)·poly(dG)·poly(dC)+ were investigated by circular and linear dichroism (CD and LD). The appearance of monomeric positive CD at a low [porphyrin]/[DNA] ratio and bisignate CD at a high ratio of the CoTMPyP-poly(dA)·poly(dT) complex is almost identical with its triplex counterpart. Similarity in the CD spectra was also observed for the CoTMPyP-poly(dG)·poly(dC) and -poly(dC)·poly(dG)·poly(dC)+ complex. This observation indicates that both monomeric binding and stacking of CoTMPyP to these polynucleotides occur at the minor groove. However, different binding geometry of CoTMPyP, when bind to AT- and GC-rich polynucleotide, was observed by LD spectrum. The difference in the binding geometry may be attributed to the difference in the interaction between polynucleotides and CoTMPyP: in the GC polynucleotide case, amine group protrude into the minor groove while it is not present in the AT polynucleotide.  相似文献   

19.
Summary The effect of Aclacinomycin B (ACM-B), an anthracycline antitumor antibiotic, on the DNA-dependent RNA synthesis using single- and double-stranded DNAs of known base content and sequence is studied. The data show that ACM-B effectively inhibits the double-stranded DNA-directed RNA synthesis with a preference of poly[d(A-T)] > poly[d(G-C)] > poly[d(I-C)]. In contrast, it has no inhibitory effect on the template function of single-stranded DNA (e.g. poly dA, poly dT, and poly dC). These results suggest that the mechanism of ACM-13 inhibition, like other anthracycline antibiotics, is by intercalation. In addition to the base specificity, there are also dramatic differences in inhibition depending on the base sequence in the DNA template. Thus, ACM-13 preferentially inhibits the alternating double-stranded copolymers over the double-stranded homopolymers; e.g. poly [d(A-T)] is inhibited to a greater extent than poly dA · poly dT and poly [d(G-C)] is inhibited more than poly dG · poly dC. Since the inhibition by ACM-13 can be totally abolished when assayed in excess amount of DNA, this result suggests that ACM-B inhibition of RNA synthesis is solely on the DNA template (which is in support of the intercalation model), and has ruled out the possibility that ACM-B may also exert an inhibitory effect on the activity of RNA polymerase per se.  相似文献   

20.
Abstract

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

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

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

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