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 396 毫秒
1.
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   

2.
We have measured the CD, isotropic absorption, and LD of poly[d(A)]–poly[d(T)] and poly[d(AT)]–poly[d(AT)] in the vacuum-uv spectral region. The reduced dichroism (LD divided by isotropic absorption) varied as a function of wavelength and was independent of shear gradient. Thus, the bases are not perpendicular to the helix axis in solution. Since the directions of the transition dipoles are known, the orientations of the bases in the polymers can be calculated from the reduced dichroism spectra. The results show that the base normals are tilted at angles greater than 25°, with respect to the helix axis, and thymine is tilted more than adenine for both polymers. The tilt axes of adenine and thymine are not parallel, indicating a large propeller twist. Space-filling models of poly[d(A)]–poly[d(T)] and poly[(AT)]–poly[d(AT)] are built based on our results, and the conformations of the two (A + T) polymers in solution are discussed.  相似文献   

3.
Psi compaction of poly[d(AT)].poly[d(AT)]   总被引:1,自引:0,他引:1  
Y A Shin  S L Feroli  G L Eichhorn 《Biopolymers》1986,25(11):2133-2148
The compaction of poly[d(A–T)] · poly[d(A–T)] by Co(III) is accompanied by the formation of ψ(+)- and ψ(-)-structures. The chirality of the ψ-structure depends on the Co(III) concentration, ionic strength, temperature, pH, and the chain length of the polymer. The two forms can be readily interconverted by manipulating these factors. Phase diagrams have been constructed that demonstrate the regions of stability of the enantiomers as a function of two variables, while other factors are held constant. At critical points in the phase diagram the two forms are in such unstable equilibrium that mechanical motion will cause ψ(+) ? ψ(-) interconversion. The formation of both ψ(+)- and ψ(-)-structures by the action of Co(III) on poly[d(A–T)] · poly[d(A–T)] contrasts markedly with the behavior of poly[d(G–C)] · poly[d(G–C)] in similar circumstances by forming only the ψ(+)-structure and that of native DNA to produce no ψ at all. Thus the base sequence is important in determining the structure of chirally associated DNA molecules.  相似文献   

4.
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   

5.
Thermodynamic parameters of melting process (DeltaHm, Tm, DeltaTm) 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 degrees C, but ZnTOEpyp(4), on the contrary, predominantly bound with AT-rich sites and increased DNA main stage Tm by 18 degrees C, and Tm of poly(dA)poly(dT) increased by 40 degrees 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 approximately 30-40 degrees C. Invariability of DeltaHm 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 DeltaTm caused by increase of added ZnTOEpyp(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 DeltaT approximately 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   

6.
Apparent molar adiabatic compressibilities and apparent molar volumes of poly[d(A-T)].poly[d(A-T)], poly(dA).poly(dT), DNA and poly(A).poly(U) in aqueous solutions were determined at 1 degree C. The change of concentration increment of the ultrasonic velocity upon replacing counter ion Cs+ by the Mg2+ ion was also determined for these polymers. The following conclusions have been made: (1) the hydration of the double helix of poly(dA).poly(dT) is remarkably larger than that of other polynucleotides; (2) the hydration of the AT pair in the B-form DNA is larger than that of the GC pair; (3) the substitution of Cs+ for Mg2+ ions as counter ions results in a decrease of hydration of the system polynucleotide plus Mg2+, and (4) the magnitude of this dehydration depends on the nucleotide sequence; the following rule is true: the lesser is a polynucleotide hydration, the larger dehydration upon changing Cs+ for Mg2+ ions in the ionic atmosphere of polynucleotide.  相似文献   

7.
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   

8.
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   

9.
W Saenger  U Heinemann 《FEBS letters》1989,257(2):223-227
In DNA oligonucleotides crystallized in the A form, the nucleotides adopt standard conformation except for steps 5'-CpG-3' where reduced base-pair twist and a sliding motion of the base pairs along their long axes causes pronounced interstrand guanine-guanine overlap. As a consequence, torsion angles alpha, beta and gamma are consistently trans, trans, trans instead of the common-gauche, trans, +gauche. This conformation significantly increases the intraresidue distance between the guanine base and the 5'-phosphate group. A molecular model of poly d(G-C).poly d(G-C) built with these structural characteristics in the A form, which we call A2-DNA, shows that rotation of the guanosine sugar into the syn orientation is easily achieved and pushes the base pair across the helix axis. If successive guanosines are changed this way, a smooth transformation occurs to the left-handed Z-DNA. We suggest that A- and A2-DNA forms of poly d(G-C).poly d(G-C) are metastable and that the actual transition is B in equilibrium (A in equilibrium A2) in equilibrium Z-DNA.  相似文献   

10.
Wrinkled DNA.   总被引:15,自引:9,他引:6       下载免费PDF全文
The B form of poly d(GC):poly d(GC) in orthorhombic microcrystallites in oriented fibers has a secondary structure in which a dinucleotide is the repeated motif rather than a mononucleotide as in standard, smooth B DNA. One set of nucleotides (probably GpC) has the same conformations as the smooth form but the alternate (CpG) nucleotides have a different conformation at C3'-O3'. This leads to a distinctive change in the orientation of the phosphate groups. Similar perturbations can be detected in other poly d(PuPy):poly d(PuPy) DNAs such as poly d(IC):poly d(IC) and poly d(AT):poly d(AT) in their D forms which have tetragonal crystal environments. This suggests that such perturbations are intrinsic to all stretches of duplex DNA where purines and pyrimidines alternate and may play a role in the detection and exploitation of such sequences by regulatory proteins.  相似文献   

11.
12.
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   

13.
D M Gray  R L Ratliff 《Biopolymers》1975,14(3):487-498
Ultraviolet circular dichroism spectra have been obtained in aqueous solutions in the presence and absence of ethanol for a synthetic DNA, poly[d(AC):d(GT)], a synthetic RNA, poly[r(AC):r(GU)], and two DNA:RNA hybrids, poly[d(AC):r(GU)] and poly[r(AC):d(GT)]. In the absence of ethanol, we find that the RNA and DNA spectra are dissimilar, while the spectra of the hybrids show differing degrees of similarity to that of the RNA. In the presence of 60–80% ethanol by weight, the spectra of the DNA and both hybrids become much closer to the spectrum of the RNA, which remains relatively unchanged. We interpret the results as indicating that DNA can undergo a change to an A-type conformation in the presence of ethanol and that the DNA:RNA hybrids are not wholly restricted to an RNA-like conformation in the absence of ethanol.  相似文献   

14.
The nuclease reactivity and specificity of a cloned tract of poly X (dA-dT) X poly(dA-dT) has been explored. Digestion with DNAse I, Mung Bean nuclease, S1 nuclease, DNAse II, and copper (1,10-phenanthroline)2 on a 256 base pair restriction fragment containing d(AT)14A revealed a dinucleotide repeat structure for the alternating sequence. Furthermore, conditions which wind or unwind the linear DNA had little effect on the reactivity of the AT insert. These preferred cleavages offer insights to structural alterations within the DNA helix which differ from A, B, or Z-DNA. Nucleation into flanking sequences by this structural alteration was not observed.  相似文献   

15.
FTIR study of netropsin binding to poly d(A-T) and poly dA.poly dT   总被引:1,自引:0,他引:1  
Complexes between netropsin and two polynucleotides containing only AT base pairs (poly d(A-T) and poly dA.poly dT) have been prepared at various drug/base pair ratios and studied in solution by Fourier Transform Infrared Spectroscopy. The drug is shown to interact in the narrow groove of poly d(A-T) with the C2O2 carbonyl of thymines and the N3 groups of adenines. Moreover the spectral modifications allow us to propose the existence of interactions at the level of the deoxyribose. No effect is detected on the phosphate groups when netropsin is progressively added. In the case of poly dA.poly dT the interaction seems much weaker as if the high propeller twist of the homopolymer would make the accessibility of the drug to the minor groove more difficult.  相似文献   

16.
Structure of the beta-form of poly d(A).poly d(U)   总被引:1,自引:0,他引:1  
The crystalline beta-form of the sodium salt of poly d(A).poly d(U) trapped in oriented fibers forms a Watson-Crick base-paired, 10(1) double-helix of pitch 3.2 nm. Two molecules are present in a monoclinic unit cell apparently isomorphous with beta-poly d(A).poly d(T). The two chains in each molecule both carry C2'-endo puckered furanose rings but are conformationally not identical. The orientations of the A:U base-pairs relative to the helix-axis are distinctly different from those in classical B-DNA and the overall morphology of the duplex in which they reside resembles that of the alpha-forms of poly (purine).poly (pyrimidine) DNA duplexes previously reported.  相似文献   

17.
An X-ray fiber diffraction study of the synthetic DNA duplex poly d(Abr5U).poly d(Abr5U) shows that its sodium salt adopts an unexceptional A-DNA-like structure. Similar to A-DNA, two molecules are packed in a monoclinic unit cell (a = 2.23 nm, b = 4.14 nm, c = 5.61 nm and alpha = beta = gamma = 90 degrees) of space group C2. Because of its dinucleotide chemical motif, the c-repeat is twice that in A-DNA but, notably, corresponding backbone conformation angles of adjacent nucleotides are almost identical. This is in marked contrast to many B-like conformations of polydinucleotides.  相似文献   

18.
The sodium dodecyl sulfate driven dissociation reactions of daunorubicin (1), mitoxantrone (2), ametantrone (3), and a related anthraquinone without hydroxyl groups on the ring or side chain (4) from calf thymus DNA, poly[d(G-C)]2, and poly[d(A-T)]2 have been investigated by stopped-flow kinetic methods. All four compounds exhibit biphasic dissociation reactions from their DNA complexes. Daunorubicin and mitoxantrone have similar dissociation rate constants that are lower than those for ametantrone and 4. The effect of temperature and ionic strength on both rate constants for each compound is similar. An analysis of the effects of salt on the two rate constants for daunorubicin and mitoxantrone suggests that both of these compounds bind to DNA through a mechanism that involves formation of an initial outside complex followed by intercalation. The daunorubicin dissociation results from both poly[d(G-C)]2 and poly[d(A-T)]2 can be fitted with a single exponential function, and the rate constants are quite close. The ametantrone and 4 polymer dissociation results can also be fitted with single exponential curves, but with these compounds the dissociation rate constants for the poly[d(G-C)]2 complexes are approximately 10 times lower than for the poly[d(A-T)]2 complexes. Mitoxantrone also has a much slower dissociation rate from poly[d(G-C)]2 than from poly[d(A-T)]2, but its dissociation from both polymers exhibits biphasic kinetics. Possible reasons for the biphasic behavior with the polymers, which is unique to mitoxantrone, are selective binding and dissociation from the alternating polymer intercalation sites and/or dual binding modes of the intercalator with both side chains in the same groove or with one side chain in each groove.  相似文献   

19.
Model sequences for evaluation of the GC dimer sequence repetition on synthesis success were prepared and analyzed by HPLC. Contiguous d(G-C) or d(C-G) sequences have a deleterious effect on DNA oligonucleotide synthesis. The critical number seems to be about 6 GCs in a row. If the GCs are separated by other nucleotides, the effect is not as severe.  相似文献   

20.
1H NMR has been used to investigate the mode of binding to d(AT)5.d(AT)5 of a series of bis(acridine) derivatives connected by different types of linker chains. The length and character (ionic, aliphatic, rigid, and flexible) of the linker chains are found to have a profound effect on the binding of these derivatives to the DNA. Bis(acridine) derivatives with linker chains shorter than 9 A monointercalate under the conditions used in the NMR study, whereas those bis(acridines) with chains of 9.8 A or longer bisintercalate. We find no evidence for the violation of the so-called neighbor exclusion principle. Although all of the bis(acridines) contain the same chromophores, their NMR spectra clearly demonstrate that they form complexes with d(AT)5.d(AT)5 which have different structures. This emphasizes the important effect that the linker chain has on the structure of the intercalation complex.  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