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 0 毫秒
1.
The molecular structure of poly (dT).poly (dA).poly (dT) has been determined and refined using the continuous x-ray intensity data on layer lines in the diffraction pattern obtained from an oriented fiber of the DNA. The final R-value for the preferred structure is 0.29 significantly lower than that for plausible alternatives. The molecule forms a 12-fold right-handed triple-helix of pitch 38.4 A and each base triplet is stabilized by a set of four Crick-Watson-Hoogsteen hydrogen bonds. The deoxyribose rings in all the three strands have C2'-endo conformations. The grooveless cylindrical shape of the triple-helix is consistent with the lack of lateral organization in the fiber.  相似文献   

2.
Conformational analysis of four stranded DNA helices poly(dT).poly(dA).poly(dA).poly(dT) with parallel arrangement of the identical sugar-phosphate chains connected by twofold symmetry has been performed. All possible models of symmetrical base binding were checked. By the potential energy optimization the dihedral angles and helices parameters of stable conformations of four stranded polynucleotides were calculated. The dependences of conformational energy on the base complex structure and mutual orientation of the poly(dA).and poly(dT) chains were studied. Possible biological functions of four stranded helices are discussed.  相似文献   

3.
Coralyne is a small crescent-shaped molecule known to intercalate duplex and triplex DNA. We report that coralyne can cause the complete and irreversible disproportionation of duplex poly(dT)·poly(dA). That is, coralyne causes the strands of duplex poly(dT)·poly(dA) to repartition into equal molar equivalents of triplex poly(dT)·poly(dA)·poly(dT) and poly(dA). Poly(dT)·poly(dA) will remain as a duplex for months after the addition of coralyne, if the sample is maintained at 4°C. However, disproportionation readily occurs upon heating above 35°C and is not reversed by subsequent cooling. A titration of poly(dT)·poly(dA) with coralyne reveals that disproportionation is favored by as little as one molar equivalent of coralyne per eight base pairs of initial duplex. We have also found that poly(dA) forms a self-structure in the presence of coralyne with a melting temperature of 47°C, for the conditions of our study. This poly(dA) self-structure binds coralyne with an affinity that is comparable with that of triplex poly(dT)·poly(dA)·poly(dT). A Job plot analysis reveals that the maximum level of poly(dA) self-structure intercalation is 0.25 coralyne molecules per adenine base. This conforms to the nearest neighbor exclusion principle for a poly(dA) duplex structure with A·A base pairs. We propose that duplex disproportionation by coralyne is promoted by both the triplex and the poly(dA) self-structure having binding constants for coralyne that are greater than that of duplex poly(dT)·poly(dA).  相似文献   

4.
The paper presents results obtained in conformational analysis of homopolymeric four-stranded poly(dT).poly(dA).poly(dA).poly(dT) DNA helices in which the pairs of strands with identical bases are parallel and have a two-fold symmetry axis. All possible models of base binding to yield a symmetric complex have been considered. The dihedral angles of sugar-phosphate backbones and helix parameters, which are consistent with the minima of conformational energy for four-stranded DNAs, have been determined using the results of optimization of conformational energy calculated at atom-atom approximation. Potential energy is shown to depend on the structure of base complexes and on the mutual orientation of unlike strands. Possible biological functions of four-stranded helices are discussed.  相似文献   

5.
Poly(dA).poly(dT), but not B-form DNA, is specifically recognized by experimentally induced anti-kinetoplast or anti-poly(dA).poly(dT) immunoglobulins. Antibody binding is completely competed by poly(dA).poly(dT) and poly(dA).poly(dU) but not by other single- or double-stranded DNA sequences in a right-handed B-form. Antibody interaction with poly(dA).poly(dT) depends on immunoglobulin concentration, incubation time and temperature, and is sensitive to elevated ionic strengths. Similar conformations, for example, (dA)4-6 X (dT)4-6, in the kinetoplast DNA of the parasite Leishmania tarentolae are also immunogenic and induce specific anti-poly(dA).poly(dT) antibodies. These antibody probes specifically recognize nuclear and kinetoplast DNA in fixed flagellated kinetoplastid cells as evidenced by immunofluorescence microscopy. Anti-poly(dA).poly(dT) immunofluorescence is DNase-sensitive and competed by poly(dA).poly(dT), but not other classical double-stranded B-DNAs. Thus, these unique cellular B'-DNA helices are immunogenic and structurally similar to synthetic poly(dA).poly(dT) helices in solution.  相似文献   

6.
R Losa  S Omari    F Thoma 《Nucleic acids research》1990,18(12):3495-3502
It was suggested that poly(dA).poly(dT) rich sequences in yeast Saccharomyces cerevisiae act as elements of constitutive promoters by exclusion of nucleosomes (Struhl, K. (1985). Proc. Natl. Acad. Sci. USA 82, 8419-8423). We have mapped the chromatin structure of the pet56-his3-ded1 region in minichromosomes and show that the poly(dA).poly(dT) sequences are located in nuclease sensitive regions. DNA fragments from the nuclease sensitive promoter region of DED1 were used for nucleosome reconstitution in vitro. We show that all sequences can form nucleosome cores and that the poly(dA).poly(dT) sequence can be incorporated in nucleosome cores. The results suggest that the nuclease sensitivity found in vivo is not established by poly(dA).poly(dT) mediated exclusion of nucleosomes.  相似文献   

7.
The effect of hydrostatic pressure upon the DNA duplex, poly(dA)poly(dT), and its component single strands, poly(dA) and poly(dT) has been studied by fourier-transform infrared spectroscopy (FT-IR). The spectral data indicate that at 28 degrees C and pressures up to 12 kbar (1200 MPa) all three polymers retain the B conformation. Pressure causes the band at 967 cm(-1), arising from water-deoxyribose interactions, to shift to higher frequencies, a result consistent with increased hydration at elevated pressures. A larger pressure-induced frequency shift in this band is observed in the single stranded polymers than in the double stranded molecule, suggesting that the effect of pressure on the hydration of single strands may be greater than upon a double stranded complex. A pressure-dependent hypochromicity in the bands attributed to base stacking indicates that pressure facilitates the base stacking in the three polymers, in agreement with previous assessments of the importance of stacking in the stabilization of DNA secondary structure at ambient and high pressures.  相似文献   

8.
The binding of propidium to poly(dA).poly(dT) [poly(dA.dT)] and to poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]] has been compared under a variety of solution conditions by viscometric titrations, binding studies, and kinetic experiments. The binding of propidium to poly[d(A-T)2] is quite similar to its binding to calf thymus deoxyribonucleic acid (DNA). The interaction with poly(dA.dT), however, is quite unusual. The viscosity of a poly(dA.dT) solution first decreases and then increases in a titration with propidium at 18 degrees C. The viscosity of poly[d(A-T)2] shows no decrease in a similar titration. Scatchard plots for the interaction of propidium with poly(dA.dT) show the classical upward curvature for positive cooperativity. The curvature decreases as the temperature is increased in binding experiments. A van't Hoff plot of the observed binding constants yields an apparent positive enthalpy of approximately +6 kcal/mol for the propidium-poly(dA.dT) interaction. Propidium binding to poly[d(A-T)2] shows no evidence for positive cooperativity, and the enthalpy change for the reaction is approximately -9 kcal/mol. Both the magnitude of the dissociation constants and the effects of ionic strength are quite similar for the dissociation of propidium from poly(dA-T)2] and from poly[d(A-T)2], suggesting that the intercalated states are similar for the two complexes. The observed association reactions, under pseudo-first-order conditions, are quite different. Plots of the observed pseudo-first-order association rate constant vs. polymer concentration have much larger slopes for propidium binding to poly[d(A-T)2] than to poly(dA.dT).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The basic assumption of Dickerson and Kopka (J. Biomole. Str. Dyns. 2, 423, 1985) that the conformation of poly(dA).poly(dT) in solution is identical to the AT rich region of the single crystal structure of the Dickerson dodecamer is not supported by any experimental data. In poly(dA).poly(dT), NOE and Raman studies indicate that the dA and dT units are conformationally equivalent and display the (anti-S-type sugar)-conformation; incorporation of this nucleotide geometry into a double helix leads to a conventional regular B-helix in which the width of the minor groove is 8A. The derived structure is consistent with all available experimental data on poly(dA).poly(dT) obtained under solution conditions. In the crystal structure of the dodecamer, the dA and dT units have distinctly different conformations-dA residues adopt (anti, S-type sugar pucker), while dT residues belong to (low anti, N-type sugar pucker). These different conformations of the dA and dT units along with the large propeller twist can be accommodated in a double helix in which the minor groove is shrunk from 8A to less than 4A. In the conventional right handed B-form of poly(dA).poly(dT) with the 8A wide minor groove, netropsin has to bind asymmetrically along the dA strand to account for the NOE and chemical shift data and to generate a stereochemically sound structure (Sarma et al, J. Biomole. Str. Dyns. 2, 1085, 1985).  相似文献   

10.
In this work, we report on the binding of the novel antitumor agent CC-1065 to poly(dA).poly(dT) and to mixtures of dA and dT oligomers as determined by electronic absorption and circular dichroism (CD) methods. In addition, the DNA binding properties of CC-1065 and its binding mechanism are compared to those of netropsin. CC-1065 binds to the polymer by at least three mechanisms to produce one irreversibly and two reversibly bound species. One reversibly bound species is moderately stable, but in time (days), it converts to the irreversibly bound species. Both of these species bind within the minor groove of the polymer and exhibit intense CC-1065 induced CD spectra. The other reversibly bound species does not acquire an induced CD. CC-1065 forces B-form duplex formation between mixtures of single strand dA and dT oligomers and binds irreversibly to the duplexes without showing the presence of an intermediate, reversibly bound species. The induced CD increases with increasing length of the oligomer, from the 5-mer (barely detectable CD) to the 14-mer (intense CD). The 7-, 10- and 14-mer mixtures bind about 1, between 1 and 2, and between 2 and 3 CC-1065 molecules, respectively. Computer graphic models of the CC-1065-DNA complex show that the covalent adduct of CC-1065 and unreacted CC-1065 can attain the same close van der Waals contacts between adenine C2 hydrogens and antibiotic CH groups that were observed in the crystal structure of the netropsin-DNA complex. These contacts may account for the dA-dT base pair binding specificity of CC-1065 and for the stability of the reversibly bound CC-1065 species.  相似文献   

11.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

12.
R W Behling  D R Kearns 《Biochemistry》1986,25(11):3335-3346
The structure of poly(dA).poly(dT) in aqueous solution has been studied by using 1H two-dimensional nuclear Overhauser effect (2D NOE) spectroscopy and relaxation rate measurements on the imino and nonexchangeable protons. The assignments of the 1H resonances are determined from the observed cross-relaxation patterns in the 2D NOE experiments. The cross-peak intensities together with the measured relaxation rates show that the purine and pyrimidine strands in poly(dA).poly(dT) are equivalent in aqueous solution. The results are consistent with a right-handed B-form helix where the sugars on both strands are in the C2'-endo/anti configuration. These observations are inconsistent with a proposed heteronomous structure for poly(dA).poly(dT) [Arnott, S., Chandrasekaran, R., Hall, I. H., & Puigjaner, L. C. (1983) Nucleic Acids Res. 11, 4141-4155]. The measured relaxation rates also show that poly(dA).poly(dT) has fast, large-amplitude local internal motions (+/- 20-25 degrees) in solution and that the amplitudes of the base and sugar motions are similar. The motion of the bases in poly(dA).poly(dT) is also similar to that previously reported for poly(dA-dT).poly(dA-dT) and poly(dG-dC).poly(dG-dC) [Assa-Munt, N., Granot, J., Behling, R. W., & Kearns, D. R. (1984) Biochemistry 23, 944-955; Mirau, P. A., Behling, R. W., & Kearns, D. R. (1985) Biochemistry 24, 6200-6211].  相似文献   

13.
14.
The conformation of DNA's with adenine-thymine tracts exhibiting retardation in electrophoretic migration and considered as curved were investigated in solution by CD and RAMAN spectroscopy. The following curved multimers with adenine tracts but of different flanking sequences d(CA5TGCC)n, d(TCTCTA6TATATA5)n, d(GA4T4C)n yield CD spectroscopic features indicating a non-B structure of the dA.dT tract with similarities to polyd(A).polyd(T). We suggest that adenine-thymine bases in these multimers contain some of the distinctive conformational features of poly(A).polyd(T) probably with large propeller twist found by NMR (Behling and Kearns, 1987) and by X-ray diffraction on oligonucleotides containing a tract of adenines (Nelson et al. 1987, Coll et al; 1987; DiGabriele et al. 1989). Some elements of distinctive CD features of the contiguous adenines run are also observed in the straight multi-9-mer d(CA5GCC)n which lacks in-phase relation to the helical repeat. Despite the presence of the TpA step in the straight multimer d(GT4A4)n, the altered dA.dT conformation is not completely destroyed. Interruption of adenine tract by a guanine in d(CAAGAATGCC)n leads to a B-like conformation and to a normal electrophoretic mobility. The Raman spectra reveal a rearrangement of the sugar-phosphate backbone of dA.dT tract in the multimer d(CA5TGCC)n with respect to that of polydA.polydT. This is reflected in the presence of an unique Raman band associated to C2'-endo sugar with a predominant contribution of C1'-exo puckering which is exhibited by the multimer whereas two distinct Raman bands characterize poly(dA).poly(dT) backbone conformation.  相似文献   

15.
B C Sang  D M Gray 《Biochemistry》1987,26(23):7210-7214
Circular dichroism (CD) data indicated that fd gene 5 protein (G5P) formed complexes with double-stranded poly(dA.dT) and poly[d(A-T).d(A-T)]. CD spectra of both polymers at wavelengths above 255 nm were altered upon protein binding. These spectral changes differed from those caused by strand separation. In addition, the tyrosyl 228-nm CD band of G5P decreased more than 65% upon binding of the protein to these double-stranded polymers. This reduction was significantly greater than that observed for binding to single-stranded poly(dA), poly(dT), and poly[d(A-T)] but was similar to that observed for binding of the protein to double-stranded RNA [Gray, C.W., Page, G.A., & Gray, D.M. (1984) J. Mol. Biol. 175, 553-559]. The decrease in melting temperature caused by the protein was twice as great for poly[d(A-T).d(A-T)] as for poly(dA.dT) in 5 mM tris(hydroxymethyl)aminomethane hydrochloride (Tris-HCl), pH 7. Upon heat denaturation of the poly(dA.dT)-G5P complex, CD spectra showed that single-stranded poly(dA) and poly(dT) formed complexes with the protein. The binding of gene 5 protein lowered the melting temperature of poly(dA.dT) by 10 degrees C in 5 mM Tris-HCl, pH 7, but after reducing the binding to the double-stranded form of the polymer by the addition of 0.1 M Na+, the melting temperature was lowered by approximately 30 degrees C. Since increasing the salt concentration decreases the affinity of G5P for the poly(dA) and poly(dT) single strands and increases the stability of the double-stranded polymer, the ability of the gene 5 protein to destabilize poly(dA.dT) appeared to be significantly affected by its binding to the double-stranded form of the polymer.  相似文献   

16.
The effect of berenil on plasmid DNA replication was studied on pBR322-derived plasmids containing poly(dA)poly(dT) sequences. In comparison to the parental plasmid pBR322, plasmid pKH47 harboring 100 bp of poly(dA)poly(dT) at the PvuII site showed a decrease in plasmid yield in the presence of berenil. This effect was also observed in pVL26, a related plasmid in which the location of the poly(dA)poly(dT) region had been shifted to the EcoRV site in pBR322. [(3)H]Thymidine incorporation experiments indicated that DNA synthesis may be affected in these plasmids in the presence of the drug. Bromodeoxyuridine incorporation experiments coupled to Cs(2)SO(4) equilibrium density gradient centrifugation indicated that the lower plasmid yield was due to an inhibition of DNA replication by berenil. We have also found that berenil induces DNA degradation in plasmids containing the homopolymer. Our studies strongly suggest that the effect of berenil on plasmid replication and DNA stability results from its binding to the poly(dA)poly(dT) region present in these plasmids. Moreover, we have found a correlation between the position of the poly(dA)poly(dT) region and this inhibitory effect. Thus, plasmid pKH47, containing the poly(dA)poly(dT) region most proximal to the origin of pBR322 replication, was most severely affected.  相似文献   

17.
X-ray diffraction in fibres revealed that the calcium salt of poly(dA).poly(dT) is a 10-fold double helix with a pitch of 3.23 nm. The opposite sugar-phosphate chains in the refined model are characterized by a complete conformational equivalence and contain sugars in a conformation close to C2'-endo. As a result a new model of the sodium salt of poly(dA).poly(dT) has been constructed, which is different from the Heteronomous DNA proposed earlier (S. Arnott et al., Nucl. Acids Res. 11, 4141 (1983)). The new model of Na-poly(dA).poly(dT) has conformationally similar opposite chains; it is a structure of the B-type, rather like that of Ca-poly(dA).poly(dT).  相似文献   

18.
A unique deletion covering around 43% of the pBR322 genome was found after attempting to insert 100 or 200 bp poly(dA).poly(dT) into the EcoRV site of pBR322 DNA. This result was not observed if an equivalent size heterologous DNA or a larger poly(dA).poly(dT) fragment of 10-20,000 bp was introduced at the same site. DNA sequencing analysis at the junctions suggests that a specific intramolecular pairing may be involved in the formation of this deletion mutant.  相似文献   

19.
A Levy  P Weisman-Shomer  M Fry 《Biochemistry》1989,28(18):7262-7267
Distamycin A, a polypeptide antibiotic, binds to dA.dT-rich regions in the minor groove of B-DNA. By virtue of its nonintercalating binding, distamycin acts as a potent inhibitor of the synthesis of DNA both in vivo and in vitro. Here we report that distamycin paradoxically stimulates Escherichia coli DNA polymerase I (pol I), its large (Klenow) fragment, and bacteriophage T4 DNA polymerase to copy oligo(dA).poly(dT) in vitro. It is found that distamycin increases the maximum velocity (Vmax) of the extension of the oligo(dA) primer by pol I without affecting the Michaelis constant (Km) of the primer. Gel electrophoresis of the extended primer indicates that the antibiotic specifically increases the rate of addition of the first three dAMP residues. Lastly, in the presence of both distamycin and the oligo(dT)-binding protein factor D, which increases the processivity of pol I, a synergistic stimulation of polymerization is attained. Taken together, these results suggest that distamycin stimulates synthesis by increasing the rate of initiation of oligo(dA) extension. The stimulatory effect of distamycin is inversely related to the stability of the primer-template complex. Thus, maximum stimulation is exerted at elevated temperatures and with shorter oligo(dA) primers. That distamycin increases the thermal stability of [32P](dA)9.poly(dT) is directly demonstrated by electrophoretic separation of the hybrid from dissociated [32P](dA)9 primer. It is proposed that by binding to the short primer-template duplex, distamycin stabilizes the oligo(dA).poly(dT) complex and, therefore, increases the rate of productive initiations of synthesis at the primer terminus.  相似文献   

20.
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