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.
N V Kumar  G Govil 《Biopolymers》1984,23(10):1995-2008
With a view to understanding the role of hydrogen bonds in the recognition of nucleic acids by proteins, hydrogen bonding between the bases and base pairs of nucleic acids and the amino acids (Asn, Gln, Asp and Glu, and charged residues Arg+, Glu?, and Asp?) has been studied by a second-order perturbation theory. Binding energies have been calculated for all possible configurations involving a pair of hydrogen bonds between the base (or base pair) and the amino acid residue. Our results show that the hydrogen bonding in these cases has a large contribution from electrostatic interaction. In general, the charged amino acids, compared to the uncharged ones, form more stable complexes with bases or base pairs. The hydrogen-bond energies are an order of magnitude smaller than the Coulombic interaction energies between basic amino acids (Lys+, Arg+, and His+) and the phosphate groups of nucleic acids. The stabilities of the complexes of amino acids Asn, Gln, Asp, and Glu with bases are in the order: G–X > C–X > A–X U–X or T–X, and G · C–X > A · T(U)–X, where X is one of these amino acid residues. It has been shown that Glu? and Asp? can recognize guanine in single-stranded nucleic acids; Arg+ can recognize G · C base pairs from A · T base pairs in double-stranded structures.  相似文献   

2.
The energies of interactions between guanine and cytosine in various mutual positions were calculated by the methods of molecular mechanics with refined atom-atom potential functions and the quantum mechanics theory of density functional. Both methods indicate three types of mutual positions of bases in local energy minima. These types correspond to (1) nearly coplanar base positions with intermolecular hydrogen bond formation (base pairing); (2) arrangements of two bases in nearly parallel planes one above another (base stacking); and (3) nearly perpendicular positions of base planes. According to the calculations, the global energy minimum corresponds to the Watson-Crick base pair with three hydrogen bonds. A specific feature of the pair is a transition from many positions of type (2) to positions of type (1) without any energy barrier. This feature is revealed by both methods. Another special feature of this pair is a deviation, for most of mutual base positions, of the amine group atoms from the ring plane, the deviation being more pronounced for Gua. These features are important for understanding the conformational behavior of DNA fragments and the RNA structure.  相似文献   

3.
Raman spectra of polyribocytidylic acid show the formation of an ordered single-stranded structure [poly(rC)] at neutral pH and an ordered double-stranded structure containing hemiprotonated bases [poly(rC)·poly(rC+)] in the range 5.5 > pH > 3.7. Below 40°C, poly(rC) contains stacked bases and a backbone geometry of the A-type, both of which are gradually eliminated by increasing the temperature to 90°C. Below 80°C, poly(rC)·poly(rC+) contains bases which are hydrogen bonded and stacked and a backbone geometry also of the A-type. In this structure the bases of each strand are shown to be structurally identical, i.e., hemiprotonated, and therefore distinct from both neutral and protonated cytosines. Infrared and Raman spectra indicate the existence of a center of symmetry with respect to the paired cytosine residues, which suggests that the additional proton per base pair is shared equally by the two hydrogen-bonded bases. Denaturation of poly(rC)·poly(rC+) occurs cooperatively (tm ≈ 80°C) with elimination of base stacking, base pairing, and the A-helix geometry. Each of the separated strands of the denatured complex is shown to contain comparable amounts of both neutral and protonated cytosines, most likely in alternating sequence [poly(rC, rC+)]. In both poly(rC, rC+) and poly(rC), at 90°C, the backbones do not exhibit the phosphodiester Raman frequencies characteristic of other disordered polyribonucleotide chains. This is interpreted to mean that the single strands, though devoid of base stacking and A-type structure, contain uniformly ordered backbones of a specific type. Fully protonated poly(rC+), on the other hand, forms no ordered structure and may be characterized as a disordered (random chain) polynucleotide at all temperatures. Several Raman lines of poly(rC) are absent from the spectrum of poly(rC)·poly(rC+) and vice versa. These frequencies, assigned mainly to vibrations of the ribose groups, suggest that the furanose ring conformations are different in the single-stranded and double-stranded structures of polyribocytidylic acid. Several other Raman group frequencies have been identified and correlated with the polymer secondary structures.  相似文献   

4.
Abstract

We have performed a conformational analysis of DNA double helices with parallel directed backbone strands connected with the second order symmetry axis being at the same time the helix axis. The calculations were made for homopolymers poly(dA) · poly(dA), poly(dC) · poly(dC), poly(dG) poly(dG), and poly(dT) · poly(dT). All possible variants of hydrogen bonding of base pairs of the same name were studied for each polymer. The maps of backbone chain geometrical existence were constructed. Conformational and helical parameters corresponding to local minima of conformational energy of “parallel” DNA helices, calculated at atom-atom approximation, were determined. The dependence of conformational energy on the base pair and on the hydrogen bond type was analysed. Two major conformational advantageous for “parallel” DNA's do not depend much on the hydrogen-bonded base pair type were indicated. One of them coincided with the conformational region typical for “antiparallel” DNA in particular for the B-form DNA Conformational energy of “parallel” DNA depends on the base pair type and for the most part is similar to the conformational energy of “antiparallel” B-DNA.  相似文献   

5.
Wilma K. Olson 《Biopolymers》1978,17(4):1015-1040
Semiempirical potential energy functional used previously to account successfully for the mean-square unperturbed dimensions and nmr coupling constants of randomly coiling polynucleotides are used, after modifications, to account for base stacking and interstrand hydrogen bonding, and to evaluate the conformational energies of single- and double-stranded polynucleotide helices. Attention is focused upon the variety of A-genus helices with local backbone conformations resembling the known double-helical structures of RNA. Distinct structural differences between single- and double-stranded helices are predicted from the energy calculations. A second point of interest is the apparent failure of two conformationally identical left-handed polynucleotide chains to form a left-handed duplex. The third major observation of the study is the wide morphological variety of theoreticaly allowed right-handed polynucleotide duplexes. In addition to the familiar double helix stabilized by horizontal base stacking and hydrogen bonding, an unusual vertical double helix is predicted to form between complementary bases fixed in the unusual but not energetically forbidden high anti glycosyl conformation. Experimental results bearing upon the theoretical predictions are discussed.  相似文献   

6.
K J Miller 《Biopolymers》1979,18(4):959-980
An algorithm is developed that enables the routine determination of backbone conformations of nucleic acids. All atomic positions including hydrogen are specified in accord with experimental bond lengths and angles but with theoretically determined conformational angles. For two Watson-Crick base pairs at a separation of 3.38 Å, and perpendicular to a common helical axis, minimum energy configurations are found for all 10 combinations at helical angles of α ~ 36°–38°, corresponding to the B-DNA structure with C(2′)-endo sugar puckers. Backbone configurations exist only within the range 35.5° ? α ? 42°, which suggests the origin of the 10-fold helix. Calculated stacking energies for the B-DNA structure increases for each of the clustered groups of base pairs: G·C with G·C, G·C with A·T, and A·T with A·T, and they are in approximate agreement with experimental observations. The counter-clockwise helix is examined, and physically meaningful structures are found only when the helical axes of successive base pairs are disjointed.  相似文献   

7.
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   

8.
The heat capacities of the single-stranded and double-stranded forms of polyadenylic acid, polyuridylic acid, and poly(uridylic and adenylic acid) were determined with a drop heat capacity calorimeter. In addition, the temperature dependence of the apparent partial heat capacity (?Cp) was measured with a newly developed differential scanning calorimeter. The calculated ΔCp at 28°C for the transition poly(A)·poly(A) ? 2 poly(A) was found to be 165 ± 24 cal/Kmol-base pair, compared with a value of 140 ± 28 for the transition poly(A)·poly(U) ? poly(A) + poly(U). The temperature dependence of ?Cp of single-stranded poly(U) was consistent with the conclusion that it is totally unstacked at temperatures above 15°C. The temperature dependence of ?Cp of single-stranded poly(A) was used to determine the base-stacking parameters for poly(A). The experimental results are consistent with a stacking enthalpy change of ?8.5 ± 0.1 kcal/mol bases and a cooperativity parameter σ of 0.57 ± 0.03 for the stacking of adenine bases. These results demonstrate that the heat capacity of single-stranded polynucleotides is greater than that of the double-stranded forms. This increased heat capacity is mainly the result of the temperature dependence of the base-stacking interactions in the single-stranded form.  相似文献   

9.
N V Kumar  G Govil 《Biopolymers》1984,23(10):2009-2024
Stacking of aromatic amino acids tryptophan (Trp), tyrosine (Tyr), phenylalanine (Phe), and histidine (His) with bases and base pairs of nucleic acids has been studied. Stacking energies of the amino acid–base (or base pair) complexes have been calculated by second-order perturbation theory. Our results show that, in general, the predominant contribution to the total stacking energy comes from the dispersion terms. In these cases, repulsion energy is greater than the sum of electrostatic and polarization energies. In contrast to this, interaction of histidine with the bases and base pairs is largely Coulombic in nature. The complexes of guanine with aromatic amino acids are more stable than the corresponding complexes of adenine. Among pyrimidines, cytosine forms the most stable complexes with the aromatic amino acids. The G · C base pair has the highest affinity with aromatic amino acids among various sets of base pairs. Optimized geometries of the stacked complexes show that the aromatic moieties overlap only partially. The heteroatom of one residue generally overlaps with the other aromatic moiety. There is a considerable degree of configurational freedom in the stacked geometries. The role of stacking in specific recognition of base sequences by proteins is discussed.  相似文献   

10.
M W Springgate  D Poland 《Biopolymers》1973,12(10):2241-2260
The cooperative nature of the binding between polycytidylate and the oligoinosinates I(pI)5–10 has been determined. Using the data of Tazawa, Tazawa, and Ts'o, it is shown that knowledge of the slope of the adsorption isothern allows one to determine the oligomer-polymer binidng constant, the oligomer–oligomer interaction constant, and the average degree of association (cooperative clustering) of the oligomers on the polymer. Knowledge of the above equilibrium constants as a function of temperature yields the respective thermodynamic parameters; no assumptions need to be made about the nature of the equilibrium constants or the thermodynamic parameters. For very long chains of polycytidylate, simple, explicit relations are given for the determination of the equilibrium constants involved. For finite chains of polycytidylate, the calculation of a single graph for each oligomer and polymer size allows the equilibrium constants to be determined for all experimental conditions of temperature and concentration. We find that the enthalpy and entropy of binding an oligomer n, bases to be δHn = ±13.7 ? n(6.65) and δSn = +32.5 ? n(18.8) given, respectively, in kcal/mole and e.u.; these parameters predict a melting temperature of 81°C for the poly(I)·poly(C) complex compared with the experimental value of 75°C. If the enthalpy is interpreted as arising from a sum of hydrogen bonding and stacking interactions, then the enthalpy of stacking is ?13.7 kcal/mole while the enthalpy of hydrogen bonding is +7 ± 4 kcal/mole; the positive enthalpy of hydrogen bonding presumably is a result of the fact that in the inosine-cytosine base pair, only two of the three sites on cytosine can hydrogen bond, the third being blocked from hydrogen bonding with water. The enthalpy of interaction between neighboring bound oligomers is found to be ?10.4 kcal/mole while the corresponding entropy is ?26.1 e.u. The binding is bound to be cooperative, though the extent of clustering varies markedly with temperature; the average number of oligomers in a cluster on the polymer is found to about five at a melting temperature of 25°C. The approach and equations given have generally applicability to oligomer-polymer associations.  相似文献   

11.
The thermodynamics of self-association (stacking) of free bases and nucleotides, intramolecular stacking in dinucleotides, nearest-neighbour base pair stacking interactions in duplex DNA and RNA, and the formation of hairpin loops illustrate enthalpy/entropy compensations. Large stacking exothermicities are associated with large negative entropy changes that ensure that delta G is small, permitting readily reversible associations in solution. We rationalise enthalpy/entropy compensations with reference to residual motions and torsional vibrations which make a larger entropic contribution to binding when - delta H approximately kT (thermal energy at room temperature), than when - delta H >> kT. We present a factorisation of experimental free energies for helix formation in terms of approximate contributions from the restriction of rotations, hydrophobic interactions, electrostatic interactions due to base stacking, and contributions from hydrogen bonding, and estimate the adverse free energy cost per rotor (mainly entropy) of ordering the phosphate backbone as between 1.9 and 5.4 kJ mol-1 [averaged over 12 rotors per base pair for A-U on A-U stacking (lower limit), and G-C on C-G stacking (upper limit)]. The largest cost is associated with the most exothermic stacking interactions, while the range of values is consistent with earlier conclusions from data on the fusion of hydrocarbon chains (lower value), and with entropy changes in covalent isomerisations of small molecules involving severe restrictions (upper value).  相似文献   

12.
Molecular models describing intermediates that may lead to proflavin-induced 1 bp deletions during in vitro polymerization by E. coli DNA polymerase I Klenow fragment are proposed. The models provide structural explanations for the fact that the induced frameshifts always occur opposite template bases that are adjacent to 5' pyrimidines and are based on the underlying hypothesis that the deletions arise because the polymerase passes by a template base without copying it. Because the most frequent mutations are opposite Pu in the template sequence 5' Py Pu 3', a single-strand loop-out model was constructed for this sequence and proflavin was added, using structures found in crystalline oligonucleotides and their complexes with proflavin. The model seeks to rationalize the roles of the 5' pyrimidine and proflavin in facilitating the bypass. Four potential roles for proflavin in mutagenesis are described: 1) stacking on the looped-out base; 2) stacking on the base pair immediately preceding the site of mutation; 3) hydrogen bonding with the 5' pyrimidine; 4) hydrogen bonding with the phosphate backbone. These models point to the possibility that a number of proflavin-DNA interactions may be involved. In contrast, modeling does not suggest a role for classically intercalated proflavin in frameshift mutagenesis arising during in vitro DNA polymerization.  相似文献   

13.
A theoretical study was performed using density functional theory (DFT) to investigate hydrogen bonding interactions in signature complexes formed between keto-9H guanine (Gua) and aspartic acid (Asp) at neutral pH. Optimized geometries, binding energies and the theoretical IR spectra of guanine, aspartic acid and their corresponding complexes (Gua-Asp) were calculated using the B3LYP method and the 6-31+G(d) basis set. Stationary points found to be at local minima on the potential energy surface were verified by second derivative harmonic vibrational frequency calculations at the same level of theory. AIM theory was used to analyze the hydrogen bonding characteristics of these DNA base complex systems. Our results show that the binding motif for the most stable complex is strikingly similar to a Watson-Crick motif observed in the guanine-cytosine base pair. We have found a range of hydrogen bonding interactions between guanine and aspartic acid in the six complexes. This was further verified by theoretical IR spectra of ω(C-H---O-H) cm−1 stretches for the Gua-Asp complexes. The electron density plot indicates strong hydrogen bonding as shown by the 2p z dominant HOMO orbital character.  相似文献   

14.
Every AE  Russu IM 《Biopolymers》2007,87(2-3):165-173
Aromatic stacking and hydrogen bonding between nucleobases are two of the key interactions responsible for stabilization of DNA double-helical structures. The present work aims at defining the specific contributions of these interactions to the stability of individual base pairs in DNA. The two DNA double helices investigated are formed, respectively, by the palindromic base sequences 5'-dCCAACGTTGG-3' and 5'-dCGCAGATCTGCG-3'. The strength of the N==H...N inter-base hydrogen bond in each base pair is characterized from the measurement of the protium-deuterium fractionation factor of the corresponding imino proton using NMR spectroscopy. The structural stability of each base pair is evaluated from the exchange rate of the imino proton, measured by NMR. The results reveal that the fractionation factors of the imino protons in the two DNA double helices investigated fall within a narrow range of values, between 0.92 and 1.0. In contrast, the free energies of structural stabilization for individual base pairs span 3.5 kcal/mol, from 5.2 to 8.7 kcal/mol (at 15 degrees C). These findings indicate that, in the two DNA double helices investigated, the strength of N==H...N inter-base hydrogen bonds does not change significantly depending on the nature or the sequence context of the base pair. Hence, the variations in structural stability detected by proton exchange do not involve changes in the strength of inter-base hydrogen bonds. Instead, the results suggest that the energetic identity of a base pair is determined by the number of inter-base hydrogen bonds, and by the stacking interactions with neighboring base pairs.  相似文献   

15.
The effects of incorporation of the modified nucleobases, 2,6-diaminopurine (D) (substituting for adenine) and 7-chloro-1,8-naphthyridin-2-(1H)-one (bicyclic thymine, bT) (substituting for thymine), that stabilize PNA·DNA duplex formation by increasing hydrogen bonding and/or base pair stacking interactions have been studied by thermal denaturation in terms of thermodynamics. Although the stabilizing effect of the bT base (in contrast to that of D base) is abolished upon addition of dimethyl formamide, thereby indicating that the stabilization is predominantly due to hydrophobic stacking forces, duplex stabilization was found to be enthalpic for both nucleobases. Increased stabilization (although not fully linearly) was observed with increasing numbers of modified bases, and single base sequence discrimination was only slightly compromised, but showed significant dependence on the sequence context.  相似文献   

16.
Abstract

Molecular models describing intermediates that may lead to proflavin-induced 1 bp deletions during in vitro polymerization by E. coli DNA polymerase I Klenow fragment are proposed. The models provide structural explanations for the fact that the induced frameshifts always occur opposite template bases that are adjacent to 5′ pyrimidines and are based on the underlying hypothesis that the deletions arise because the polymerase passes by a template base without copying it. Because the most frequent mutations are opposite Pu in the template sequence 5′ Py Pu 3′, a single-strand loop-out model was constructed for this sequence and proflavin was added, using structures found in crystalline oligonucleotides and their complexes with proflavin. The model seeks to rationalize the roles of the 5′ pyrimidine and proflavin in facilitating the bypass. Four potential roles for proflavin in mutagenesis are described: 1) stacking on the looped-out base; 2) stacking on the base pair immediately preceding the site of mutation; 3) hydrogen bonding with the 5′ pyrimidine; 4) hydrogen bonding with the phosphate backbone. These models point to the possibility that a number of proflavin-DNA interactions may be involved. In contrast, modeling does not suggest a role for classically intercalated proflavin in frameshift mutagenesis arising during in vitro DNA polymerization.  相似文献   

17.
V N Damle 《Biopolymers》1972,11(9):1789-1816
Interaction of poly U (polyuridylic acid) and adenosine is studied by following the changes in ultraviolet absorbance in the wavelength region near the isochromic wave-length for the complex formation. The interaction is studied as a function of temperature, concentration of adenosine, and ionic strength, while the concentration of poly U was held constant. It is confirmed that only the three-stranded complex with the stoichiometry 1A to 2U is formed and that it dissociates directly into free poly U and adenosine. No discontinuity of any kind was apparent in the melting curves, and poly U was found to possess no ordered structure above 10°C under the conditions used. The results were, therefore, analyzed in terms of an exact helix–coil equilibrium theory using the mismatching model, i.e., assuming that either completely formed base triplet or completely free unbonded bases only exist, and that the two sections of the polymer chains forming closed loops need not contain the same number of unbonded bases. Self-association of free adenosine was taken into consideration. (Base triplet is analog of base pair for a three-stranded helical complex. It refers to a unit of three coplanar bases, in this case two uracils and one adenine, hydrogen bonded to one another to form a triplet. Such triplets may stack over one another along the helical axis, and when they are so stacked the bases of two triplets next to each other may have stacking interactions between them.) The values for enthalpy and entropy changes, both per mole of base triplets, were obtained for the following processes at neutral pH and moderate to high salt concentrations. (1) Growfh: Binding of one adenosine molecule to two uracil residues (one from each poly U strand) to form a base triplet next to an already formed base triplet with which it has stacking interactions, a process that involves both hydrogen bonding and base stacking interactions, ΔHs, = ?19 ± 2 kcal, ΔSs = ?55 ± 6 clausius; (2) Initiation: Binding of one adenosine molecule to two uracil residues (one from each poly U strand) to form an isolated base triplet, a process that involves only hydrogen bonding interactions, ΔHb* = 4.5 ± 2 kcal, ΔSb* = 6.6 ± 3 clausius; and (3)Interruption: Unstacking of two stacked base triplets initially next to each other by formation of an interruption (viz. a closed loop) between them, a process that involves only base stacking interactions, ΔHb = 23.5 ± 3 kcal, ΔSb = 61.6 ± 7 clausius, where the entropy changes include contributions other than the configurational entropy of closed loops. The discrepancy between our results and the calorimetric ΔHs of ?13 kcal is attributed to (i) the possible effects of salt arid polymer on the self-association of free adenosine, (ii) the uncertainty in the value of the parameter for the probability of ring closure, and (iii) the contributions due to the partial molal enthalpy of the solvent and the unstacking of any poly U structure to the calorimetric enthalpy.  相似文献   

18.
We have performed a conformational analysis of DNA double helices with parallel directed backbone strands connected with the second order symmetry axis being at the same time the helix axis. The calculations were made for homopolymers poly(dA).poly(dA), poly(dC).poly(dC), poly(dG) poly(dG), and poly(dT).poly(dT). All possible variants of hydrogen bonding of base pairs of the same name were studied for each polymer. The maps of backbone chain geometrical existence were constructed. Conformational and helical parameters corresponding to local minima of conformational energy of "parallel" DNA helices, calculated at atom-atom approximation, were determined. The dependence of conformational energy on the base pair and on the hydrogen bond type was analysed. Two major conformational advantageous for "parallel" DNA's do not depend much on the hydrogen-bonded base pair type were indicated. One of them coincided with the conformational region typical for "antiparallel" DNA, in particular for the B-form DNA. Conformational energy of "parallel" DNA depends on the base pair type and for the most part is similar to the conformational energy of "antiparallel" B-DNA.  相似文献   

19.
Stacking interaction between the aromatic heterocyclic bases plays an important role in the double helical structures of nucleic acids. Considering the base as rigid body, there are total of 18 degrees of freedom of a dinucleotide step. Some of these parameters show sequence preferences, indicating that the detailed atomic interactions are important in the stacking. Large variants of non‐canonical base pairs have been seen in the crystallographic structures of RNA. However, their stacking preferences are not thoroughly deciphered yet from experimental results. The current theoretical approaches use either the rigid body degrees of freedom where the atomic information are lost or computationally expensive all atom simulations. We have used a hybrid simulation approach incorporating Monte‐Carlo Metropolis sampling in the hyperspace of 18 stacking parameters where the interaction energies using AMBER‐parm99bsc0 and CHARMM‐36 force‐fields were calculated from atomic positions. We have also performed stacking energy calculations for structures from Monte‐Carlo ensemble by Dispersion corrected density functional theory. The available experimental data with Watson–Crick base pairs are compared to establish the validity of the method. Stacking interaction involving A:U and G:C base pairs with non‐canonical G:U base pairs also were calculated and showed that these structures were also sequence dependent. This approach could be useful to generate multiscale modeling of nucleic acids in terms of coarse‐grained parameters where the atomic interactions are preserved. This method would also be useful to predict structure and dynamics of different base pair steps containing non Watson–Crick base pairs, as found often in the non‐coding RNA structures. © 2015 Wiley Periodicals, Inc. Biopolymers 105: 212–226, 2016.  相似文献   

20.
Abstract

The crystal structures of five double helical DNA fragments containing non-Watson-Crick complementary base pairs are reviewed. They comprise four fragments containing G·T base pairs: two deoxyoctamers d(GGGGCTCC) and d(GGGGTCCC) which crystallise as A type helices; a deoxydodecamer d(CGCGAATTTGCG) which crystallises in the B-DNA conformation; and the deoxyhexamer d(TGCGCG), which crystallises as a Z-DNA helix. In all four duplexes the G and T bases form wobble base pairs, with bases in the major tautomer forms and hydrogen bonds linking N1 of G with 02 of T and 06 of G with N3 of T. The X-ray analyses establish that the G·T wobble base pair can be accommodated in the A, B or Z double helix with minimal distortion of the global conformation. There are, however, changes in base stacking in the neighbourhood of the mismatched bases. The fifth structure, d(CGCGAATTAGCG), contains the purine purine mismatch G·A where G is in the anti and A in the syn conformation. The results represent the first direct structure determinations of base pair mismatches in DNA fragments and are discussed in relation to the fidelity of replication and mismatch recognition.  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