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.
Jean Sturm 《Biopolymers》1982,21(6):1189-1206
A temperature-jump relaxation study of the interaction of tilorone with different polynucleotides and DNA has been performed. A single relaxation time, attributed to the intercalation step, has been observed in the case of poly[d(A-T)]·poly[d(A-T)], poly[d(A-C)]·poly[d(G-T)], poly[d(G-C)]·poly[d(G-C)], and poly(dG)·poly(dC). No intercalation into poly(dA)·poly(dT) occurs, and the interaction with poly(dG)·poly(dC) is different from what is observed with the other intercalating homopolymers. Refinement of the binding model is suggested from the analysis of the kinetic data. The relaxation curves obtained with DNA are well simulated based on a binding mechanism where DNA is considered a heterogeneous lattice and each type of site behaves as if it were located in the corresponding homopolymer. Poly(dA)·poly(dT) shows a unique behavior: studies of the effects of concentration and temperature indicate that tilorone acts as a probe of a process involving the polynucleotide alone. This process appears to be related to the dynamic structure of the nucleic acid and is detectable only when the bound dye is not intercalated.  相似文献   

2.
The rate constants of 1H----3H exchange between water and C8H-groups of purinic residues of alternating polynucleotides: poly[d(A-T)].poly[d(A-T)] (I), poly[d(G-C)].poly[d(G-C)] (II), poly[d(A-C)].poly[d(G-T)] (III) and homopolynucleotides: poly(dA).poly(dt) (IV), poly(dG).poly(dC) (V), as well as DNA E. coli, was determined in 0.15 M NaCl at 25 degrees C. The retardation of exchange observed at these conditions (compared to that of the B-form DNA) is in agreement with the model of B-alternating structure for the (I) and is attributed to the co-existence of B- and A-conformers for the (V) in solution. Absence of distinguishable differences in exchange rate constants for purinic residues of the (II), (III) and (IV) (compared to that of the B-form DNA) evidences that conformations of these polynucleotides in solution are similar to "canonical" B-form DNA and don't correlate with the model of "heteronomous" DNA which was proposed for (IV).  相似文献   

3.
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   

4.
We have tested 21 different basis sets of synthetic DNA circular dichroism spectra and have slected one for use in spectral analyses of natural DNAs. This “standard” set consists of spectra of eight polymers: poly[d(A-A-T)·d(A-T-T)], poly[d(A-G-G)·d(C-C-T)], poly[d(A-T)·d(A-T)], poly[d(G-C)·d(G-C)], poly[d(A-G)·d(C-T)], poly[d(A-C)·d(G-T)], poly[d(A-T-C)·d(G-A-T)], and poly[d(A-G-C)·d(G-C-T)]. This basis set, applied according to the first-neighbor polymer procedure of Gray and Tinoco, allows a more uniformly accurate spectral analysis of six natural complex DNAs and eight (A+T)-rich satellite DNAs for base composition and first-neighbor frequencies than was previously possible. We find that spectra of poly[d(A)·d(T)] and/or poly[d(A-C-T-)·d(A-G-T)] are not generally required for good analysis results but we show in this and the following paper that these spectra are needed for the most accurate analyses of some satellite DNAs.  相似文献   

5.
The nature of binding of Ru(phen) 2+ (I), Ru(bipy) 2+ (II), Ru(terpy) 2+ (III) (phen = 1,10-phenanthroline, bipy 3 = 2,2'-bipyridyl, 3 terpy = 2,2'2," - 2 terpyridyl) to DNA, poly[d(G-C)] and poly[d(A-T)] has been compared by absorption, fluorescence, DNA melting and DNA unwinding techniques. I binds intercalatively to DNA in low ionic strength solutions. Topoisomerisation shows that it unwinds DNA by 22 degrees +/- 1 per residue and that it thermally stabilizes poly[d(A-T)] in a manner closely resembling ethidium. Poly[d(A-T)] induces greater spectral changes on I than poly[d(G-C)] and a preference for A-T rich regions is indicated. I binding is very sensitive to Mg2+ concentration. In contrast to I the binding of II and III appears to be mainly electrostatic in nature, and causes no unwinding. There is no evidence for the binding of the neutral Ru(phen)2 (CN)2 or Ru(bipy)2 (CN)2 complexes. DNA is cleaved, upon visible irradiation of aerated solutions, in the presence of either I or II.  相似文献   

6.
Sequence-dependent variation in the conformation of DNA   总被引:50,自引:0,他引:50  
The specificity of action of the enzyme DNAase I on double-stranded DNA polymers of defined sequence has been investigated. The results obtained with the alternating copolymers poly[d(A-T)] · poly[d(A-T)] and poly[d(G-C)] · poly[d(G-C)] support the suggestion of Klug et al. (1979) that regions of double-stranded DNA containing alternating purine-pyrimidine sequences may exist as structural variants of the classical B-form under physiological salt conditions. Digestion of defined oligomers containing alternating dG-dC sequences indicate that these too exist in some “alternating-B” structure in solution under similar conditions. The results obtained with the oligomers also provide a number of insights into the mode of action of DNAase I.In the case of the B-DNA dodecamer d(C-G-C-G-A-A-T-T-C-G-C-G), for which the crystal structure has been solved (Dickerson &; Drew, 1981), there is a very good correlation between the sites of rapid DNAase I cutting and positions of high local helical twist.  相似文献   

7.
The effect of Hg2+ and Ag+ on the buoyant density (rho) of four synthetic DNA polymers, poly[d(A-T)]; poly(dA) - poly(dT); oikt[d(G-C)]; and poly(dG) - poly(dC), was investigated. The buoyant density of poly[d(A-T)] in Cs2SO4 increased dramatically after complexing with Hg2+, but little change in the buoyant density of other polymers resulted except at very high molar ratios of Hg2+/DNA-P (rf). Hg2+ raised the thermal transition temperature (Tm) of alternating polymers and lowered the Tm of homopolymers. Measurements in the preparative ultracentrifuge indicated that lowered Tm correlated with Hg2+-induced strand separation of one homopolymer [poly(dA) - poly(dT)], but strand separation was not observed with another homopolymer [poly(dG) - poly(dC)] complexed with Hg2+. When Ag+ was mixed with the polymers, the buoyant density of poly(dG) - poly(dC) increased most markedly. A substantial increase in the buoyant density of poly[d(A-T)] and a small increase in the buoyant density of poly[d(G-C)] were also observed. The Tm changes induced by Ag+ were not related in any obvious way to buoyant density changes. These findings indicate that nucleotide sequence as well as overall base composition is of importance in understanding the buoyant density changes induced by metal ions. Although these data do not allow construction of a detailed molecular model of polymer-metal ion interactions, they may be used to explain much of the behavior of naturally occurring DNA sequences, such as heterochromatic satellite sequences and 5 S and rRNA sequences, in Hg2+/Cs2SO4 and Ag+/Cs2SO4 gradients.  相似文献   

8.
Raman spectra of six synthetic polydeoxyribonucleotide duplexes with different base sequences have been examined in aqueous solutions with different salt or nucleotide concentrations. Detailed conformational differences have been indicated between B and Z forms of poly[d(G-C)] X poly[d(G-C)], between B forms of poly[d(G-C)] X poly[d(G-C)] and poly[d(G-m5C)] X poly[d(G-m5C)], between A and B forms of poly(dG) X poly(dC), between B and "CsF" forms of poly[d(A-T)] X poly[d(A-T)], between B forms of poly[d(A-U)] X poly[d(A-U)] and poly[d(A-T)] X poly[d(A-T)], and between low- and high-salt (CsF) forms of poly(dA) X poly(dT). The Raman spectrum of calf-thymus DNA in aqueous solution was also observed and was compared with the Raman spectra of its fibers in A, B, and C forms.  相似文献   

9.
The interactions of two positional isomers and one analogue of meso-tetra (4-N-methylpyridyl) porphine, with the synthetic polynucleotides poly[d(A-T)] . poly[d(A-T)] and poly[d(G-C)] . poly[d(G-C)] have been investigated by circular dichroism. All four porphyrins were found to bind to the polynucleotides as shown by the induction of circular dichroism in their Soret bands. Furthermore, the sign of the induced ellipticity reflects selective occupation of binding sites by the porphyrin ligands. The conformational lability of poly[d(A-T)] X poly[d(A-T)] was found to be appreciable as micromolar amounts of meso-substituted 4-N-methylpyridyl, 3-N-methylpyridyl, and p-N-trimethylanilinium porphines induced a CD spectrum similar but not identical to that of DNA in the Z-form, i.e. a negative band at 280 nm and a positive band at 259 nm. The effect of porphyrin binding to poly[d(G-C)] X poly[d(G-C)] was less pronounced and dissimilar to that seen in the AT polymer.  相似文献   

10.
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   

11.
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   

12.
We examined the binding geometry of Co-meso-tetrakis (N-methyl pyridinium-4-yl)porphyrin, Co-meso-tetrakis (N-n-butyl pyridinium-4-yl)porphyrin and their metal-free ligands to poly[d(A-T)(2)] and poly[d(G-C)(2)] by optical spectroscopic methods including absorption, circular and linear dichroism spectroscopy, and fluorescence energy transfer technique. Signs of an induced CD spectrum in the Soret band depend only on the nature of the DNA sequence; all porphyrins exhibit negative CD when bound to poly[d(G-C)(2)] and positive when bound to poly[d(A-T)(2)]. Close analysis of the linear dichroism result reveals that all porphyrins exhibit outside binding when complexed with poly[d(A-T)(2)], regardless of the existence of a central metal and side chain. However, in the case of poly[d(G-C)(2)], we observed intercalative binding mode for two nonmetalloporphyrins and an outside binding mode for metalloporphyrins. The nature of the outside binding modes of the porphyrins, when complexed with poly[d(A-T)(2)] and poly[d(G-C)(2)], are quite different. We also demonstrate that an energy transfer from the excited nucleo-bases to porphyrins can occur for metalloporphyrins.  相似文献   

13.
Pressure-jump study of the kinetics of ethidium bromide binding to DNA   总被引:4,自引:0,他引:4  
Pressure-jump chemical relaxation has been used to investigate the kinetics of ethidium bromide binding to the synthetic double-stranded polymers poly[d(G-C)] and poly[d(A-T)] in 0.1 M NaCl, 10 mM tris(hydroxymethyl)aminomethane hydrochloride, and 1 mM ethylenediaminetetraacetic acid, pH 7.2, at 24 degrees C. The progress of the reaction was followed by monitoring the fluorescence of the intercalated ethidium at wavelengths greater than 610 nm upon excitation at 545 nm. The concentration of DNA was varied from 1 to 45 microM and the ethidium bromide concentration from 0.5 to 25 microM. The data for both polymers were consistent with a single-step bimolecular association of ethidium bromide with a DNA binding site. The necessity of a proper definition of the ethidium bromide binding site is discussed: it is shown that an account of the statistically excluded binding phenomenon must be included in any adequate representation of the kinetic data. For poly[d(A-T)], the bimolecular association rate constant is k1 = 17 X 10(6) M-1 s-1, and the dissociation rate constant is k-1 = 10 s-1; in the case of poly[d(G-C)], k1 = 13 X 10(6) M-1 s-1, and k-1 = 30 s-1. From the analysis of the kinetic amplitudes, the molar volume change, delta V0, of the intercalation was calculated. In the case of poly[d(A-T)], delta V0 = -15 mL/mol, and for poly[d(G-C)], delta V0 = -9 mL/mol; that is, for both polymers, intercalation is favored as the pressure is increased.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
L A Marky  R B Macgregor 《Biochemistry》1990,29(20):4805-4811
We report differences in the interaction of two structurally similar phenanthroline intercalators, ethidium and propidium, with poly(dA).poly(dT) and poly[d(A-T)] as a function of ionic strength based on titration microcalorimetry, fluorescence titration, and hydrostatic pressure measurements. Both ethidium and propidium bind more strongly to poly[d(A-T)].poly[d(A-T)] than to poly(dA).poly(dT). Ethidium intercalation into the latter polymer displays titrations with positive cooperativity; this is not found with propidium. The enthalpy of intercalation (delta H degrees) is exothermic for both dyes with poly[d(A-T)].poly[d(A-T)]; however, the value of this parameter is nearly zero in the case of poly(dA).poly(dT). The molar volume change (delta V degrees) accompanying dye intercalation is negative under all conditions for poly[d(A-T)].poly[d(A-T)] whereas it is positive for poly(dA).poly(dT). The changes observed in delta V degrees correlate well with the entropy changes derived from the titration and calorimetric data for this reaction. The results, interpreted in terms of the relative hydration of these two polymers, are consistent with a higher extent of hydration of poly(dA).poly(dT) relative to poly[d(A-T)].poly[d(A-T)].  相似文献   

15.
G H Shimer  A R Wolfe  T Meehan 《Biochemistry》1988,27(20):7960-7966
We have investigated the equilibrium binding of racemic 7r,8t,9t,10c-tetrahydroxy-7,8,9,10-tetrahydrobenzo[a]pyrene to the double-stranded, synthetic polynucleotides poly[d(A-T)], poly[d(G-C)], and poly[d(G-m5C)] at low binding ratios. Difference absorption spectroscopy shows a 10-nm red shift for binding to poly[d(A-T)] and an 11-nm red shift for binding to either poly[d(G-C)] or poly[d(G-m5C)]. The value of delta epsilon for binding is approximately the same for all three hydrocarbon-polynucleotide complexes. Binding of this neutral polycyclic aromatic hydrocarbon derivative to these polynucleotides is dependent upon ionic strength and temperature. Analysis of complex formation employing polyelectrolyte theory shows a greater release of counterions associated with binding to poly[d(A-T)] than with the other two polynucleotides (0.5 and ca. 0.36, respectively). Thus, sequence-selective binding of this hydrocarbon in DNA would be expected to change depending on salt concentration. The temperature dependence of binding was studied at 100 mM Na+ where the equilibrium binding constants for poly[d(A-T)] and poly[d(G-m5C)] are roughly equivalent and 6-fold greater than the binding affinity for poly[d(G-C)]. The binding to poly[d(A-T)] and poly[d(G-C)] is characterized by a delta H omicron = -7.0 kcal/mol, and the large difference in affinity constants arises from differences in negative entropic contributions. Formation of hydrocarbon-poly[d(G-m5C)] complexes is accompanied by a delta H = -9.1 kcal/mol. However, the affinity for poly[d-(G-m5C)] is the same as that for poly[d(A-T)] due to the much more negative entropy associated with binding to poly[d(G-m5C)].  相似文献   

16.
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   

17.
The binding of 9-hydroxyellipticine to calf thymus DNA, poly[d(A-T)]2, and poly-[d(G-C)]2 has been studied in detail by means of CD, linear dichroism, resonance light scattering, and molecular dynamics. The transition moment polarizations of 9-hydroxyelliptiycine were determined in polyvinyl alcohol stretched film. Spectroscopic solution studies of the DNA/drug complex are combined with theoretical CD calculations using the final 50 ps of a series of molecular dynamics simulations as input. The spectroscopic data shows 9-hydroxyellipticine to adopt two main binding modes, one intercalative and the other a stacked binding mode involving the formation of drug oligomers in the DNA major groove. Analysis of the intercalated binding mode in poly[d(A-T)]2 suggests the 9-hydroxyellipticine hydroxyl group lies in the minor groove and hydrogen bonds to water with the pyridine ring protruding into the major groove. The stacked binding mode was examined using resonance light scattering and it was concluded that the drug was forming small oligomer stacks rather than extended aggregates. Reduced linear dichroism measurements suggested a binding geometry that precluded a minor groove binding mode where the plane of the drug makes a 45° angle with the plane of the bases. Thus it was concluded that the drug stacks in the major groove. No obvious differences in the mode of binding of 9-hydroxyellipticine were observed between different DNA sequences; however, the stacked binding mode appeared to be more favorable for calf thymus DNA and poly[d(G-C)]2 than for poly[d(A-T)]2, an observation that could be explained by the slightly greater steric hindrance of the poly[d(A-T)]2 major groove. A strong concentration dependence was observed for the two binding modes where intercalation is favored at very low drug load, with stacking interactions becoming more prominent as the drug concentration is increased. Even at DNA : drug mixing ratios of 70:1 the stacked binding mode was still important for GC-rich DNAs. © 1998 John Wiley & Sons, Inc. Biopoly 46: 127–143, 1998  相似文献   

18.
A new purification technique for ‘single-stranded DNA-binding proteins’ from calf thymus permits the demonstration of a considerable heterogeneity within these proteins. Several molecular species are obtained with Mr between 24·103 and 30·103 and pI values between 6 and 8, showing significant differences with regard to the following functional properties: (1) strength of binding to single-stranded DNA; (2) lowering of melting temperature of poly[d(A-T)]; (3) stimulation of DNA polymerase α on a poly[d(A-T)] template. Analysis of trypsin digestion products demonstrates that the different molecular species share extensive primary sequence homology. Experiments with antibodies show that the different molecular species are antigenically related and that a 31 kDa protein present in low amounts in our preparations is very cross-reactive.  相似文献   

19.
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   

20.
R S Johnson 《Biochemistry》1991,30(1):198-206
A derivative of RNA polymerase containing approximately 2 pyrene equiv per enzyme molecule has been used to study the interaction of RNA polymerase with poly[d(A-T)].poly[d(A-T)] and poly[d-(G-C)].poly[d(G-C)]. As monitored by fluorescence spectroscopy, pyrenyl RNA polymerase displays a unique set of conformational changes with each synthetic polynucleotide as a function of temperature. An increase in the fluorescence intensity was observed for both polynucleotides at 5 degrees C. A decrease was observed in the case of poly[d(A-T)].poly[d(A-T)] at 25 and 37 degrees C, whereas no discernible perturbation was observed in the case of poly[d(G-C)].poly[d(G-C)]. Different salt dependencies were observed for the interaction of pyrenyl RNA polymerase with these polynucleotides at 5 and 25 degrees C. Further characterization of these interactions as well as correlation of the observed fluorescence changes to the corresponding open and closed complexes was carried out with heparin. The interaction between pyrenyl RNA polymerase and poly[d-(A-T)].poly[d(A-T)] at 25 degrees C was quantified by using two different methods.(ABSTRACT TRUNCATED AT 250 WORDS)  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