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 15 毫秒
1.
The hydrogen-deuterium exchange of AMP, uridine, poly(rA), and poly(rA) · poly(rU) was investigated by a spectral difference method using stopped-flow spectrophotometry. Proton exchange rates were measured as a function of pH, added catalysts, temperature and salt concentration. The results confirm and extend previous conclusions on the H-exchange chemistry of the bases, on the large equilibrium opening of the double helix, and on its slow opening and closing rates, but an alternative conformation for the major open state is considered. Two H-exchange rate classes are found in poly(rA) · poly(rU). The slower class represents the two exocyclic amino protons of A which exchange through a pre-equilibrium opening mechanism, therefore revealing the fraction of time the helix is open. Base-pairs are open 5% of the time at 25 °C. The faster class is assigned to the U-N-3 H proton, the rate of which is limited by helix opening. Both opening and reclosing of the duplex are slow, 2 s?1 and 40 s?1, respectively, at 25 °C. Thermodynamic parameters for the equilibrium helix opening and for the rate of opening were determined. These properties may be consistent with a simple opening involving swinging out of the U base while retaining A more or less stacked within the duplex. The results demonstrate that no faster or more populated helix-open state occurs (when structure is stable). It appears that, unlike opening—closing reactions at a helix end or a helix-coil boundary, internal base opening and closing are innately slow. One implication of this is that any chemical or biological process requiring access to sequences in the interior of a closed stable DNA duplex may be constrained to proceed only on a time scale of seconds, and not in milliseconds or microseconds.  相似文献   

2.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly (rI) · poly (rC), single-stranded poly(rC) and poly(rI), inosine, and cytosine- 5′-phosphoric acid have been examined, at various temperatures in the range 20 °C to 52 °C, by stopped-flow ultraviolet spectrophotometry, in the region 270 to 300 nm. For the solution of double-helical poly(rI) · poly(rC), two first-order deuteration reactions were found: a fast one and a slow one. At 25 °C and at pH 7.0, the rate constant was 12.3 s?1 for the fast reaction, and 0.13 s?1 for the slow reaction. The rate constant of the fast reaction is nearly equal to that of the single-stranded poly(rC) (12.6 s?1), and is assigned to the deuteration at the amino hydrogen (that is, free from the C · I hydrogen bond) of the cytosine residue. The slow reaction is attributable to the deuteration of the two hydrogens: the amino hydrogen of rC and imide hydrogen of rI, which are rapidly exchanging with each other within every rC · rI base-pair. From the observed temperature effect on this slow reaction rate, it has been concluded that there are two types of “opening process” that are relevant to the hydrogen exchange reaction; one of them is predominent in the range 47 °C to 52 °C and the other in the temperature region lower than 47 °C. The enthalpy (H) and entropy (S) differences of the “open” and “closed” forms in the former type process are ΔH = 167 kcal per mole and ΔS = 507 e.u., while in the latter ΔH = 8.1 kcal per mole and ΔS = 10 e.u..  相似文献   

3.
Hydrogen-exchange studies of I · C and G · C double helices were carried out to test the generality of conclusions reached previously in studies of adenine-containing polymers (preceding paper). The cytosine amino group shows hydrogen-exchange behavior similar to the analogous group in adenine; a pH-independent pathway and a parallel general catalysis pathway require prior separation of the base-pair and pre-equilibrium protonation at the ring N. The cytosine amino group does, however, display greater sensitivity to specific and to general catalysis than found for adenine. In the G · C helix, the ring NH proton of guanine exchanges at the opening-limited rate, as does the analogous proton in A · U and A · T pairs, while the guanine amino protons exchange without a prior opening of structure. From the observed exchange rates and the known chemistry for the pH-independent reaction, one can calculate equilibrium opening constants of 4 × 10−3 for poly(rI) · poly(rC) and perhaps one tenth of that for poly(rG) · poly(rC). Also the opening rate constant for the G · C helix is 0.01 s−1.These results, when applied to published exchange curves for DNA, indicate an equilibrium opening constant of 0.005, an opening rate constant of 0.04 s−1, and a closing rate constant of 10 s−1. (All values refer to studies at 0 °C.) These values point to the same kind of traveling-loop model for base-pair opening discussed previously for the opening reactions in adenine-containing double helices.  相似文献   

4.
Using (1)H NMR spectroscopy, the base-pair opening dynamics of an antiparallel foldback DNA triplex and the corresponding duplex has been characterized via catalyzed imino proton exchange. The triplex system was found to be in an equilibrium between a duplex and a triplex form. The exchange rate between the two forms (i.e., the on/off-rate of the third strand) was measured to be 5 s(-1) at 1 degrees C, and the base-pair dynamics of both forms were investigated separately. Both Watson-Crick and reverse Hoogsteen base pairs were found to have base-pair lifetimes in the order of milliseconds. The stability of the Watson-Crick base pairs was, however, substantially increased in the presence of the third strand. In the DNA triplex, the opening dynamics of the reverse Hoogsteen base pairs was significantly faster than the dynamics of the Watson-Crick pairs. We were able to conclude that, for both Watson-Crick and reverse Hoogsteen base pairs, spontaneous and individual opening from within the closed base triplet is the dominating opening pathway.  相似文献   

5.
The slow kinetics of annealing processes in multistranded nucleic acids is spectrophotometrically investigated using poly(A)·2poly(U) as a model system. The absorbance changes at specific wavelengths show that double-helical (A·U) base pairs appear as transient intermediates. The annealing process is identified by the enlargement of triple-helical sequences at the cost of (A·U) base pairs and unpaired (U) residues. A large time range in the reorganization of mismatched chain configurations is characterized by a logarithmic dependence on time. This observation is quantitatively described by a kinetic model developed by Jackson. In Jackson's model the rate-limiting process in the slow annealing phase of maximizing triple-helical sequences, is the removal of strand entanglements, knots, and hairpin loops by complete unwinding of those helical stretches which stabilize the mismatched configurations. The results of the present study are briefly discussed in terms of optimum conditions for hybridization experiments and for the preparation of polynucleotide complexes commonly used to produce interferons.  相似文献   

6.
The vibrational spectra and structure of poly(rA-rU)-poly(rA-rU)   总被引:2,自引:0,他引:2  
Infrared and Raman spectra of aqueous poly(rA-rU)·poly(rA-rU), the double-helical complex containing strands of alternating riboadenylate and ribouridylate residues, display significant differences from one another and from corresponding spectra of poly(rA)·poly(rU), the double-helical complex of riboadenylate and ribouridylate homopolymers. Parallel studies on the copolymer and homopolymer complexes by cesium sulfate density gradient centrifugation, ultraviolet absorption spectroscopy, hydrogenion titration, 1-N oxidation of adenine residues by monoperphthalic acid and X-ray diffraction reveal, however, that the geometry of base pairing between adenine and uracil is closely similar in each complex and apparently of the Watson-Crick type. Therefore the differences observed between vibrational spectra of poly (rA-rU)·poly (rA-rU) and poly(rA)·poly(rU) are not due to different base-pairing schemes but may be attributed to differences in vibrational coupling between vertically stacked bases. Vibrational coupling may also account for the differences between infrared and Raman spectra of the same complex. Thus, the present results indicate that infrared and Raman frequencies of RNA in the region 1750–1550 cm?1 should be dependent on the base sequence.  相似文献   

7.
The thermal stability of a new polynucleotide complex has been used to establish the hydrogen-bonding structure of three-stranded C-G·CH+ helices. In the Hoogsteen structure, the 8NH2 group of 8NH2GMP can form a third hydrogen bond to the CH+ strand, but in the alternative structure, the 8NH2 group can form no interbase hydrogen bonds. For the new complex, 8NH2GMP·2 poly(C), a transition temperature of 80°C is observed under conditions in which the corresponding complex formed with 5′-GMP has a Tm of 20°C. We conclude from this 60° elevation of transition temperature that a third hydrogen bond is formed by the 8NH2 group and that the structure must have Hoogsteen bonding. In order to be compatible with this structure in regular helices formed by U,C copolymers, A·2U bonding would also have to have a Hoogsteen structure.  相似文献   

8.
Abstract

Phase transitions were studied of the sodium salt of poly(rA) ?poly(rU) induced by elevated temperature without Ni2+ and with Ni2+ in 0.07 M concentration in D2O (~0.4 [Ni]/[P]). The temperature was varied from 20° C to 90° C. The double-stranded conformation of poly(rA)?poly(rU) was observed at room temperature (20° C—23° C) with and without Ni2+ ions. In the absence of Ni2+ ions, partial double- to triple-strand transition of poly(rA) ?poly(rU) occurred at 58° C, whereas only single-stranded molecules existed at 70° C. While poly(rU) did not display significant helical structure, poly(rA) still maintained some helicity at this temperature. Ni2+ ions significantly stabilized the triple-helical structure. The temperature range of the stable triple-helix was between 45° C and 70° C with maximum stability around 53° C. Triple-to single-stranded transition of poly(rA) ?poly(rU) occurred around 72° C with loss of base stacking in single-stranded molecules. Stacked or aggregated structures of poly(rA) formed around 86° C. Hysteresis took place in the presence of Ni2+ during the reverse transition from the triple-stranded to the double-stranded form upon cooling. Reverse Hoogsteen type of hydrogen-bonding of the third strand in the triplex was suggested to be the most probable model for the triple-helical structure. VCD spectroscopy demonstrated significant advantages over infrared absorption or the related electronic CD spectroscopy.  相似文献   

9.
P Bendel 《Biopolymers》1987,26(4):573-590
The results of semiselective 1H-nmr inversion recovery experiments on sonicated calf thymus DNA fragments are reported. The measurements were conducted in aqueous solutions containing 85% D2O, in order to reduce the dipolar contribution to the observed relaxation rates. In solutions containing 0.2M NaCl, 0.4 mM EDTA, and 10 mM cacodylate at pH = 7.0, the exchange rates of the imino protons in A-T base pairs confirm values published earlier in the literature, extrapolating to 0.25 s?1 at 25°C. Corresponding values for the G-C base pairs are published for the first time, and are about sixfold slower. The addition of up to 0.1M Tris buffer (pH = 7.3 at 25°C), caused a striking increase in the measured exchange rates for both the A-T and G-C imino protons, resembling the effect recently observed for poly(rA)-poly(rU) and poly(rI)-poly(rC), and suggesting that the exchange rates measured for nucleic acid duplexes in low buffer concentrations at neutral pH do not reflect base-pair opening rates as assumed in the past. Lower limits to the base-pair opening rates could be estimated from extrapolation of the experimental data to infinite buffer concentration, and are 1 × 103 s?1 for the A-T, and 50 s?1 for the G-C, base paris at 62°C.  相似文献   

10.
Abstract

The nucleic acid triplexes poly d(T)·poly d(A)·poly d(T), poly (U)·poly (A)·poly (U), and poly (I)·poly (A)·poly (I) display a sort of continuity between each other. However, their morphologies present their own individuality which, considering those of their parent duplexes, are quite unexpected. This comparison helps to understand triplex structure-function relationship. While helical parameters are functions of the sugar pucker, low values of WC and Hoogsteen base-pair propellers is commonplace for triplexes and the Hoogsteen base-pair geometry monitors the effects of the interstrand phosphates charge-charge repulsion.

Synopsis

The nucleic acid triplexes poly d(T)·poly d(A)·poly d(T), poly(U)·poly(A)·poly(U), and poly (I)·poly (A)·poly (I) present distinct morphologies. Considering those of their parent duplexes, they are also quite unexpected.  相似文献   

11.
The binding of an antitumour drug with bisquarternary ammonium heterocyclic structure, NSC-101327, to nucleic acids has been examined by using ultraviolet absorption and CD measurements. Like the minor groove-binding oligopeptides, netropsin and distamycin A, the optically inactive chromophoric system of NSC-101327 shows induced Cotton effects in the CD spectra of complexes with various DNAs, RNA and single-stranded polynucleotides. This property directly reflects interaction of NSC-101327 with different types of nucleic acids at moderate ionic strength, which contrasts with previous findings of a higher selective binding of netropsin to B-DNA. However, an efficient interactin of NSC-101327 with dA·dT basepair sequences is demonstrated by a large melting temperature increase of dA·dT-rich DNAs. NSC-101327 also reacts with dG·dC base pairs of B-DNA and forms a complex with Z-DNA of poly(br8dG-dC)·poly(br8DG-dC). The affinity of NSC-101327 to poly(dG-dC)·poly(dG-dC) is, however, lower, and the CD spectral binding effect depends on the ionic strength. The CD results of the complex with poly(dA-dT)·poly(dA-dT) suggests at least two binding modes, in accordance with previous conclusions. This is indicated by a clear-cut initial increase of the CD signal and a subsequent large decrease to negative CD signals. Competition experiments with netropsin suggest that binding of NSC-101327 occurs preferentially in the minor groove without intercalation. NSC-101327 also tends to interact with lower binding affinity to dG-dC pairs in B-DNA, with rA·rU pairs of RNA and with single-stranded polynucleotides. Thus our results suggest that NSC-101327 represents a DNA groove-binding ligand of lower basepair specificity and lower conformational selectivity compared to the B-specific netropsin probe.  相似文献   

12.
S1 is an acidic protein associated with the 3′ end of 16S RNA; it is indispensable for ribosomal binding of natural mRNA. We find that S1 unfolds single stranded stacked or helical polynucleotides (poly rA, poly rC, poly rU). It prevents the formation of poly (rA + rU) and poly (rI + rC) duplexes at 10–25 mM NaCl but not at 50–100 mM NaCl. Partial, salt reversible denaturation is also seen with coliphage MS2 RNA, E. coli rRNA and tRNA. Generally, only duplex structures with a Tm greater than about 55° are formed in the presence of S1. The protein unfolds single stranded DNA but not poly d(A·T).  相似文献   

13.
Abstract

Using Raman spectroscopy, we examined the ribose-phosphate backbone conformation, the hydrogen bonding interactions, and the stacking of the bases of the poly(U)·poly(A) ·poly(U) triple helix. We compared the Raman spectra of poly(U)·poly(A)·poly(U) in H2O and D2O with those obtained for single-stranded poly(A) and poly(U) and for double-stranded poly(A)·poly(U). The presence of a Raman band at 863 cm?1 indicated that the backbone conformations of the two poly(U) chains are different in the triple helix. The sugar conformation of the poly(U) chain held to the poly(A) by Watson-Crick base pairing is C3′ endo; that of the second poly(U) chain may be C2′ endo. Raman hypochromism of the bands associated with base vibrations demonstrated that uracil residues stack to the same extent in double helical poly(A)·poly(U) and in the triple-stranded structure. An increase in the Raman hypochromism of the bands associated with adenine bases indicated that the stacking of adenine residues is greater in the triple helix than in the double helical form. Our data further suggest that the environment of the carbonyls of the uracil residues is different for the different strands.  相似文献   

14.
The kinetics of interaction between proflavine and poly I.poly C at 25°C, neutral pH, and moderate ionic strength have been studied by relaxation methods. The qualitative features of this system resemble those previously reported by Crothers and co-workers for proflavine–DNA and proflavine–poly A·poly U interactions–two relaxations are observed coresponding to a fast bimolecular step followed by a slower isomerization. These results can best be accommodated by a two-step mechanism leading from the free dye through an “outside-bound” complex to the intercalated complex. Quantitative comparison of the various rate constants for proflavine binding to different double-helical polynucleotides shows that the rates are slower for both ribohomopolymer pairs than for DNA. The rates for poly I·poly C are approximately three times faster than these for poly A·poly U.  相似文献   

15.
Abstract

A single-point substitution of the O4′ oxygen by a CH2 group at the sugar residue of A 6 (i.e. 2′-deoxyaristeromycin moiety) in a self-complementary DNA duplex, 5′- d(C1G2C3G4A5A6T7T8C9G10C11G12)2 ?3, has been shown to steer the fully Watson-Crick basepaired DNA duplex (1A), akin to the native counterpart, to a doubly A 6:T7 Hoogsteen basepaired (1B) B-type DNA duplex, resulting in a dynamic equilibrium of (1A)→←(1B): Keq = k1/k-1 = 0.56±0.08. The dynamic conversion of the fully Watson-Crick basepaired (1A) to the partly Hoogsteen basepaired (1B) structure is marginally kinetically and thermodynamically disfavoured [k1 (298K) = 3.9± 0.8 sec?1; δH°? = 164±14 kJ/mol;-TδS°? (298K) = ?92 kJ/mol giving a δG298°? of 72 kJ/mol. Ea (k1) = 167±14 kJ/mol] compared to the reverse conversion of the Hoogsteen (1B) to the Watson-Crick (1A) structure [k-1 (298K) = 7.0±0.6 sec-1, δH°? = 153±13 kJ/mol;-TδS°? (298K) = ?82 kJ/mol giving a δG298°? of 71 kJ/mol. Ea (k-1) = 155±13 kJ/mol]. A comparison of δG298°? of the forward (k1) and backward (k-1) conversions, (1A)→←(1B), shows that there is ca 1 kJ/mol preference for the Watson-Crick (1A) over the double Hoogsteen basepaired (1B) DNA duplex, thus giving an equilibrium ratio of almost 2:1 in favour of the fully Watson-Crick basepaired duplex. The chemical environments of the two interconverting DNA duplexes are very different as evident from their widely separated sets of chemical shifts connected by temperature-dependent exchange peaks in the NOESY and ROESY spectra. The fully Watson-Crick basepaired structure (1A) is based on a total of 127 intra, 97 inter and 17 cross-strand distance constraints per strand, whereas the double A 6:T7 Hoogsteen basepaired (1B) structure is based on 114 intra, 92 inter and 15 cross-strand distance constraints, giving an average of 22 and 20 NOE distance constraints per residue and strand, respectively. In addition, 55 NMR-derived backbone dihedral constraints per strand were used for both structures. The main effect of the Hoogsteen basepairs in (1B) on the overall structure is a narrowing of the minor groove and a corresponding widening of the major groove. The Hoogsteen basepairing at the central A 6:T7 basepairs in (1B) has enforced a syn conformation on the glycosyl torsion of the 2′- deoxyaristeromycin moiety, A 6, as a result of substitution of the endocyclic 4′-oxygen in the natural sugar with a methylene group in A 6. A comparison of the Watson-Crick basepaired duplex (1A) to the Hoogsteen basepaired duplex (1B) shows that only a few changes, mainly in α, σ and γ torsions, in the sugar-phosphate backbone seem to be necessary to accommodate the Hoogsteen basepair.  相似文献   

16.
Non-canonical base pairs play important roles in organizing the complex three-dimensional folding of RNA. Here, we outline methodology developed both to analyze the spatial patterns of interacting base pairs in known RNA structures and to reconstruct models from the collective experimental information. We focus attention on the structural context and deformability of the seven pairing patterns found in greatest abundance in the helical segments in a set of well-resolved crystal structures, including (i–ii) the canonical A·U and G·C Watson–Crick base pairs, (iii) the G·U wobble pair, (iv) the sheared G·A pair, (v) the A·U Hoogsteen pair, (vi) the U·U wobble pair, and (vii) the G·A Watson–Crick-like pair. The non-canonical pairs stand out from the canonical associations in terms of apparent deformability, spanning a broader range of conformational states as measured by the six rigid-body parameters used to describe the spatial arrangements of the interacting bases, the root-mean-square deviations of the base-pair atoms, and the fluctuations in hydrogen-bonding geometry. The deformabilties, the modes of base-pair deformation, and the preferred sites of occurrence depend on sequence. We also characterize the positioning and overlap of the base pairs with respect to the base pairs that stack immediately above and below them in double-helical fragments. We incorporate the observed positions of the bases, base pairs, and intervening phosphorus atoms in models to predict the effects of the non-canonical interactions on overall helical structure.  相似文献   

17.
The interaction of Na ions with synthetic polynucleotides   总被引:1,自引:0,他引:1  
The interaction of Na+ with poly A, poly U, poly A·poly U, and Poly A·2 poly U has been investigated by means of potentiometry, by means of potentiometry, by means of a linked-function analysis of its effect on the binding of Mg++ ions, and of K+ by means of an analysis of its effect on the sedimentation coefficients of the polymers. The last method was found to be inapplicable. The results of the other two methods were found to be consistent, except in the case of poly A where the existence of base stacking, influenced by the binding of Mg++, significantly affects the linked-function analysis. The results are also consistent with the effects of the concentration of Na+ ions on the thermally induced conformational transitions of poly A·poly U and poly A·2 poly U, and with the extents of “binding” of Na+ to DNA measured by equilibrium and by transport methods. The interaction of Na+ with polynucleotides appears to be physically quite specific, although its thermodynamic basis is not clear. The extent of binding of Na+, Ψ, was found to be independent of the total Na+ concentration but a quadratic function of the extent of Mg++ binding, θ. In the absence of Mg++, Ψ = 0.35–0.38 for poly U, 0.40 for poly A, 0.59 for poly A·poly U, and 0.66 for poly A·2 poly U.  相似文献   

18.
Raman studies of nucleic acids. VII. Poly A-poly U and poly G-poly C   总被引:6,自引:0,他引:6  
L Lafleur  J Rice  G J Thomas 《Biopolymers》1972,11(12):2423-2437
Laser-excited Raman spectra of the double-helical complexes poly A·poly U and poly G·poly C are reported for 2H2O and H2O solutions. The spectra are discussed in relation to their use as quantitative reference spectra for determining the dependence of the Raman scattering of RNA on secondary structure. The Raman line at 815 cm?1, due to the phosphodiester group, exhibits the same intrinsic intensity in spectra of poly A·poly U and poly G·poly C and is thus dependent only upon the amount of ordering of the helix and not on the kinds of nucleotides involved. The hypochromic Raman lines in spectra of poly A·poly U are identified and their intensity changes are determined quantitatively over the temperature range 32–85°C. Comparison of the spectra in the 1500–1750 cm?1 region reveals that the Raman lines from carbonyl group vibrations of uracil are about sevenfold more intense than those of guanine and cytosine for both paired and unpaired states and will thus dominate the spectra of RNA. The Raman frequencies in this region are also compared with previously reported infrared frequencies and give evidence of being strongly perturbed by base-stacking interactions in the helices.  相似文献   

19.
Abstract

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

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

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

20.
P A Mirau  D R Kearns 《Biopolymers》1985,24(4):711-724
1H-nmr relaxation has been used to study the effect of sequence and conformation on imino proton exchange in adenine–thymine (A · T) and adenine–uracil (A · U) containing DNA and RNA duplexes. At low temperature, relaxation is caused by dipolar interactions between the imino and the adenine amino and AH2 protons, and at higher temperature, by exchange with the solvent protons. Although room temperature exchange rates vary between 3 and 12s?1, the exchange activation energies (Eα) are insensitive to changes in the duplex sequence (alternating vs homopolymer duplexes), the conformation (B-form DNA vs A-form RNA), and the identity of the pyrimidine base (thymine vs uracil). The average value of the activation energy for the five duplexes studied, poly[d(A-T)], poly[d(A) · d(T)], poly[d(A-U)], Poly[d(A) · d(U)], and poly[r(A) · r(U)], was 16.8 ± 1.3 kcal/mol. In addition, we find that the average Eα for the A.T base pairs in a 43-base-pair restriction fragment is 16.4 ± 1.0 kcal/mol. This result is to be contrasted with the observation that the Eα of cytosine-containing duplexes depends on the sequence, conformation, and substituent groups on the purine and pyrimidine bases. Taken together, the data indicate that there is a common low-energy pathway for the escape of the thymine (uracil) imino protons from the double helix. The absolute values of the exchange rates in the simple sequence polymers are typically 3–10 times faster than in DNAs containing both A · T and G · C base pairs.  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