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 46 毫秒
1.
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   

2.
Flow linear dichroism is used to measure specific inclinations for each of the four bases in poly[d(AC)]·;poly[d(GT)] and poly[d(AG)]·poly[d(CT)] in both the B and A forms. For the B form in solution the bases are found to have a sizable inclination. Inclination is increased in the A form, as expected. In all cases the pyrimidines are more inclined than the purines. © 1993 John Wiley & Sons, Inc.  相似文献   

3.
Raman spectroscopy of Z-form poly[d(A-T)].poly[d(A-T)   总被引:3,自引:0,他引:3  
Helical structures of double-stranded poly[d(A-T)] in solution have been studied by Raman spectroscopy. While the classical right-handed conformation B-type spectra are obtained in the case of sodium chloride solutions, a Z-form Raman spectrum is observed by addition of nickel ions at high sodium concentration, conditions in which the inversion of the circular dichroic spectrum of poly[d(A-T)] is detected, similar to that observed for high-salt poly[d(G-C)] solutions [Bourtayre, P., Liquier, J., Pizzorni, L., & Taillandier, E. (1987) J. Biomol. Struct. Dyn. 5, 97-104]. The characterization of the Z-form spectrum of poly[d(A-T)] is proposed by comparison with previously obtained characteristic Raman lines of Z-form poly[d(G-C)] and poly[d(A-C)].poly[d(G-T)] solutions and of d(CG)3 and d(CGCATGCG) crystals [Thamann, T. J., Lord, R. C., Wang, A. H.-J., & Rich, A. (1981) Nucleic Acids Res. 9, 5443-5457; Benevides, J. M., Wang, A. H.-J., van der Marel, G. A., van Boom, J. H., Rich, A., & Thomas, G. J., Jr. (1984) Nucleic Acids Res. 14, 5913-5925]. Detailed spectroscopic data are presented reflecting the reorientation of the purine-deoxyribose entities (C2'-endo/anti----C3'-endo/syn), the modification of the phosphodiester chain, and the adenosine lines in the 1300-cm-1 region. The role played by the hydrated nickel ions in the B----Z transition is discussed.  相似文献   

4.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

5.
Formation of psi (+) and psi (-) DNA   总被引:4,自引:0,他引:4  
Y A Shin  G L Eichhorn 《Biopolymers》1984,23(2):325-335
DNA molecules can be organized into ordered aggregates of opposite handedness by complexation with polylysine and other polypeptides; we have investigated the conditions under which ψ(+) and ψ(?) structures are produced with double-helical synthetic polynucleotides. Both poly(dGdC)·poly(dGdC) and poly(dAdT)·poly(dAdT) readily form ψ(?) structures with polylysine, although the method of preparation can alter the CD spectra. The GC copolymer, which is more susceptible to conversion into A or Z conformers, forms ψ(+) structures with lysine–alanine copolypeptides more readily than the AT copolymer. Nucleotide base modifications that favor the Z structure, such as bromination and methylation, also favor ψ(+) formation, and the Co(NH3)6Cl3 reagent that readily induces the Z structure also leads to ψ(+). Thus, the production of the ψ(+) structure seems to be frequently correlated with susceptibility to A or Z formation, although there are some cases in which the B conformer also leads to ψ(+). Polyethylene glycol generally produces a ψ(?) structure; the differentiation between ψ(+) and ψ(?) structures seems to require electrically charged polymers.  相似文献   

6.
The alpha-form of poly[d(A)].poly[d(T)], observed in fibers at high (greater than 80%) relative humidity, is a 10-fold double-helical structure of pitch 3.2 nm. This new X-ray analysis shows that the two strands of the double helix are of the same kind conformationally and both B-like in containing C-2'-endo-puckered deoxyribose rings. Nevertheless, the two strands are different enough for the overall morphology of the duplex to resemble that of the heteromerous model for the drier (beta) form of poly[d(A)].poly[d(T)] in which one strand has C-2'-endo rings and the other C-3'-endo. Since the orientations of the bases in poly[d(A)].poly[d(T)] are persistently different from those of classical B-DNA it is likely that there will be local bending (about 10 degrees) at the junctions between general sequence tracts and the oligo[d(A)].oligo[d(T)] tracts that occur in some native DNAs. The conclusions about the structure of alpha-poly[d(A)].poly[d(T)] are reinforced by independent analyses of similar X-ray diffraction patterns from poly[d(A)].poly[d(U)] and poly[d(A-I)].poly[d(C-T)].  相似文献   

7.
S P Edmondson  W C Johnson 《Biopolymers》1986,25(12):2335-2348
We have measured the CD, isotropic absorption, and linear dichroism (LD) in the vacuum-uv spectral region for the B-conformations of poly[d(G)]-poly[d(C)] and poly[d(GC)]-poly[d(GC)], and for the Z-conformation of poly[d(GC)]-poly[d(GC)] formed in 70% trifluoroethanol. The reduced dichroism (LD divided by isotropic absorption) for all conformations varied with wavelength, indicating that the bases are not perpendicular to the helix axis. Since the directions of the transition dipoles are known, the inclinations and axes of inclination of each base can be determined from the wavelength dependence of the reduced dichroism spectra. The results indicate that the base normals of the (G + C) polymers in the B- and Z-conformations are tilted at angles greater than 19° with respect to the helix axis. The guanine and cytosine bases have different inclinations, and the tilt axes are not parallel. Therefore, the bases for all the (G + C) polymer conformations studied are buckled and propeller twisted.  相似文献   

8.
S P Edmondson 《Biopolymers》1987,26(11):1941-1956
Model helices for poly[d(A)]-poly[d(T)] in solution were constructed based on the base orientations determined by linear dichroism [LD; S. P. Edmondson & W. C. Johnson, Jr. (1985) Biopolymers 24 , 824–841] and refined by molecular mechanical energy minimization. The results demonstrate that poly[d(A)]-poly[d(T)] can form energetically stable conformations with the base orientations measured by LD. Further, only one of the four possible base-pair orientations that are consistent with the LD results is feasible. Models with negative base tilts had large potential energies, indicating that the LD results do not reflect base motions. The LD models of poly[d(A)]-poly[d(T)] are stabilized mainly by intrastrand base-stacking interactions, particularly for the adenine strand.  相似文献   

9.
Infrared dichroism measurements of oriented films of poly(dA)·poly(dT) and poly[d(A-T)]·poly[d(A-T)] have been made under the conditions of low salts content and high humidity for which the geometry is known. The angles which the transition moments make with the helix axis are compared with the orientations of the corresponding bonds. Except for the in-plane base model of poly[(A-T)]·poly[d(A-T)], there is no agreement. This may imply either that a model which assumes bonds and transition moments to be colinear is not acceptable or that x-ray data are inaccurate. These possibilities are discussed especially with respect to phosphate group orientation. An appendix gives the derivations of dichroic-ratio expressions for helical molecules of different symmetry types.  相似文献   

10.
We have measured the circular dichroism (CD) and absorption properties of poly[r(G-U)] and poly [d(G-T)] over a wide range of Na+ concentrations and temperatures. We find evidence for self-complexed forms of these polymers at lower temperatures and/or higher Na+ concentrations than generally needed for double-strand formation in other DNA and RNA polymers. These self-complexes could be composed of double-stranded regions with weak G·U or G·T base pairs.  相似文献   

11.
The binding of polyamines, including spermidine ( 1 ) and spermine ( 2 ), to poly[d(G-C) · d(G-C) ] was probed using spectroscopic studies of anthracene-9-carbonyl-N1-spermine ( 3 ); data from normal absorption, linear dichroism (LD), and circular dichroism (CD) are reported. Ligand LD and CD for transitions located in the DNA region of the spectrum were used. The data show that 3 binds to DNA in a manner characteristic of both its amine and polycyclic aromatic parts. With poly [(dG-dC) · (dG-dC)], binding modes are occupied sequentially and different modes correspond to different structural perturbations of the DNA. The most stable binding mode for 3 with poly[d(G-C) · d(G-C)] has a site size of 6 ± 1 bases, and an equilibrium binding constant of (2.2 ± 1.1) × 107 M?1 with the anthracene moiety intercalated. It dominates the spectra from mixing ratios of approximately 133:1 until 6:1 DNA phosphate: 3 is reached. The analogous data for poly [d(A-T) · d(A-T)] between mixing ratios 36:1 and 7:1 indicates a site size of 8.3 ± 1.1 bases and an equilibrium binding constant of (6.6 ± 3.3) × 105 M?1. Thus, 3 binds preferentially to poly [d(G-C) · d(G-C)] at these concentrations. © 1994 John Wiley & Sons, Inc.  相似文献   

12.
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   

13.
Metal-induced sequential transitions among DNA conformations   总被引:1,自引:0,他引:1  
The action of [Co(NH3)6]Cl3 on poly(dGdC) · poly(dGdC) can lead to a series of consecutive reactions, in which B-DNA is first converted to Z-DNA [M. Behe and G. Felsenfeld (1981) Proc. Natl. Acad. Sci USA 78 , 1619–1623], which in turn is transformed into an unidentified conformer that we tentatively call “U”, and finally the highly associated anisotropic ψ-DNA is produced. Conditions are given under which the sequence B ? Z ? “U” ψ(+) can be stopped at any point in the direction from left to right. The reverse processes, from right to left, occur when ψ(+) or “U”-DNA are treated with various amounts of salt concentrations and lowering temperature. Thus it is demonstrated that four conformers of poly(dGdC) · poly(dGdC) are readily interconvertible, and that Z-DNA and “U” conformers are intermediates in the reversible transformations of B- and ψ-DNA.  相似文献   

14.
A refinement of the recent results for poly[d(A)].poly[d(T)] (Alexeev et al., J. Biomol. Struct. Dyn. 4,989 (1987)) involving additional parameters of the base-pair structure and of the sugar-phosphate backbone expands the conformational potential of this polynucleotide of the B type to include the possibility of bifurcated hydrogen bonds of the kind recently discovered in crystalline deoxyoligonucleotide with lone d(A)n.d(T)n stretch (Nelson et al., Nature 330, 221 (1987)). Still, analysis of the available data and energy calculations do not seem to indicate that the bifurcated H-bonds are a crucial factor responsible for the anomalous structure of the d(A)n.d(T)n sequence. The unique structural properties of poly[d(A)].poly[d(T)] can hardly be explained without taking into account its interactions with the double-layer hydration spine in the minor groove. In view of the hydration mechanism stabilizing poly[d(A)].poly[d(T)] and of the polynucleotide's heteronomous prehistory (Arnott et al., Nucleic Acids Res. 11,4141 (1983)) we suggest that this B-type structure be called Bh.  相似文献   

15.
gp32 I is a protein with a molecular weight of 27 000. It is obtained by limited hydrolysis of T4 gene 32 coded protein, which is one of the DNA melting proteins. gp32 I itself appears to be also a melting protein. It denatures poly[d(A-T)].poly[d(A-T)] and T4 DNA at temperatures far (50-60 degrees C) below their regular melting temperatures. Under similar conditions gp32 I will denature poly[d(A-T).poly[d(A-T)] at temperatures approximately 12 degrees C lower than those measured for the intact gp32 denaturation. For T4 DNA gp32 shows no melting behavior while gp32 I shows considerable denaturation (i.e., hyperchromicity) even at 1 degree C. In this paper the denaturation of poly[d(A-T)].poly[d(A-T)] and T4 DNA by gp32 I is studied by means of circular dichroism. It appears that gp32 I forms a complex with poly[d(A-T)]. The conformation of the polynucleotide in the complex is equal to that of one strand of the double-stranded polymer in 6 M LiCl. In the gp32 I DNA complex formed upon denaturation of T4 DNA, the single-stranded DNA molecule has the same conformation as one strand of the double-strand T4 DNA molecule in the C-DNA conformation.  相似文献   

16.
Synthetic RNA poly[r(A-T)] has been synthesized and its CD spectral properties compared to those of poly[r(A-U)], poly[d(A-T)], and poly[d(A-U)] in various salt and ethanolic solutions. The CD spectra of poly[r(A-T)] in an aqueous buffer and of poly[d(A-T)] in 70.8% v/v ethanol are very similar, suggesting that they both adopt the same A conformation. On the other hand, the CD spectra of poly[r(A-T)] and of poly[r(A-U)] differ in aqueous, and even more so in ethanolic, solutions. We have recently observed a two-state salt-induced isomerization of poly[r(A-U)] into chiral condensates, perhaps of Z-RNA [M. Vorlícková, J. Kypr, and T. M. Jovin, (1988) Biopolymers 27, 351-354]. It is shown here that poly[r(A-T)] does not undergo this isomerization. Both the changes in secondary structure and tendency to aggregation are different for poly[r(A-T)] and poly[r(A-U)] in aqueous salt solutions. In most cases, the CD spectrum of poly[r(A-U)] shows little modification of its CD spectrum unless the polymer denatures or aggregates, whereas poly[r(A-T)] displays noncooperative alterations in its CD spectrum and a reduced tendency to aggregation. At high NaCl concentrations, poly[r(A-T)] and poly[r(A-U)] condense into psi(-) and psi(+) structures, respectively, indicating that the type of aggregation is dictated by the polynucleotide chemical structure and the corresponding differences in conformational properties.  相似文献   

17.
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   

18.
In previous work, it was shown that poly [d(AC) · d(GT)] could be forced into the Z form by strong dehydrating conditions, provided EDTA was not present. Presumably multivalent impurities were also necessary for the transition. In order to gain control over the B to Z transition for this DNA, we carefully removed all divalent contaminants from the sample and asked the obvious question: What ions are necessary for the transition under dehydrating conditions? We systematically investigated the effect of various multivalent ions. The common contaminants Ca2+, Mg2+, and Fe3+ will not cause the transition, but Co2+ and Ni2+ facilitate the transition, undoubtedly because of their well-known propensity to bind to purine N7. Since the transition also depends on the synergistic dehydrating action of sodium perchlorate and ethanol, we include CD spectra for the independent variations of these two factors. In addition, vacuum-uv CD spectra for the A form and various B forms of poly [d (AC) · d (GT)] are presented for the first time.  相似文献   

19.
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   

20.
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   

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