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.
Nearest-neighbor thermodynamic parameters of the ‘universal pairing base’ deoxyinosine were determined for the pairs I·C, I·A, I·T, I·G and I·I adjacent to G·C and A·T pairs. Ultraviolet absorbance melting curves were measured and non-linear regression performed on 84 oligonucleotide duplexes with 9 or 12 bp lengths. These data were combined with data for 13 inosine containing duplexes from the literature. Multiple linear regression was used to solve for the 32 nearest-neighbor unknowns. The parameters predict the Tm for all sequences within 1.2°C on average. The general trend in decreasing stability is I·C > I·A > I·T ≈ I· G > I·I. The stability trend for the base pair 5′ of the I·X pair is G·C > C·G > A·T > T·A. The stability trend for the base pair 3′ of I·X is the same. These trends indicate a complex interplay between H-bonding, nearest-neighbor stacking, and mismatch geometry. A survey of 14 tandem inosine pairs and 8 tandem self-complementary inosine pairs is also provided. These results may be used in the design of degenerate PCR primers and for degenerate microarray probes.  相似文献   

2.
The thermodynamics of base pairing is of fundamental importance. Fluorinated base analogs are valuable tools for investigating pairing interactions. To understand the influence of direct base–base interactions in relation to the role of water, pairing free energies between natural nucleobases and fluorinated analogs are estimated by potential of mean force calculations. Compared to pairing of AU and GC, pairing involving fluorinated analogs is unfavorable by 0.5–1.0 kcal mol−1. Decomposing the pairing free energies into enthalpic and entropic contributions reveals fundamental differences for Watson–Crick pairs compared to pairs involving fluorinated analogs. These differences originate from direct base–base interactions and contributions of water. Pairing free energies of fluorinated base analogs with natural bases are less unfavorable by 0.5–1.0 kcal mol−1 compared to non-fluorinated analogs. This is attributed to stabilizing C–FH–N dipolar interactions and stronger NH–C hydrogen bonds, demonstrating direct and indirect influences of fluorine. 7-methyl-7H-purine and its 9-deaza analog (Z) have been suggested as members of a new class of non-fluorinated base analogs. Z is found to be the least destabilizing universal base in the context of RNA known to date. This is the first experimental evidence for nitrogen-containing heterocylces as bioisosteres of aromatic rings bearing fluorine atoms.  相似文献   

3.
G·U wobble base pairs are the most common and highly conserved non-Watson–Crick base pairs in RNA. Previous surface maps imply uniformly negative electrostatic potential at the major groove of G·U wobble base pairs embedded in RNA helices, suitable for entrapment of cationic ligands. In this work, we have used a Poisson–Boltzmann approach to gain a more detailed and accurate characterization of the electrostatic profile. We found that the major groove edge of an isolated G·U wobble displays distinctly enhanced negativity compared with standard GC or AU base pairs; however, in the context of different helical motifs, the electrostatic pattern varies. G·U wobbles with distinct widening have similar major groove electrostatic potentials to their canonical counterparts, whereas those with minimal widening exhibit significantly enhanced electronegativity, ranging from 0.8 to 2.5kT/e, depending upon structural features. We propose that the negativity at the major groove of G·U wobble base pairs is determined by the combined effect of the base atoms and the sugar-phosphate backbone, which is impacted by stacking pattern and groove width as a result of base sequence. These findings are significant in that they provide predictive power with respect to which G·U sites in RNA are most likely to bind cationic ligands.  相似文献   

4.
Coralyne is an alkaloid drug that binds homo-adenine DNA (and RNA) oligonucleotides more tightly than it does Watson–Crick DNA. Hud’s laboratory has shown that poly(dA) in the presence of coralyne forms an anti-parallel duplex, however attempts to determine the structure by NMR spectroscopy and X-ray crystallography have been unsuccessful. Assuming adenine–adenine hydrogen bonding between the two poly(dA) strands, we constructed 40 hypothetical homo-(dA) anti-parallel duplexes and docked coralyne into the six most favorable duplex structures. The two most stable structures had trans glycosidic bonds, but distinct pairing geometries, i.e. either Watson–Crick Hoogsteen (transWH) or Watson–Crick Watson–Crick (transWW) with stability of transWH > transWW. To narrow down the possibilities, 7-deaza adenine base substitutions (dA→7) were engineered into homo-(dA) sequences. These substitutions significantly reduced the thermal stability of the coralyne-induced homo-(dA) structure. These experiments strongly suggest the involvement of N7 in the coralyne-induced A·A base pairs. Moreover, due to the differential effect on melting as a function of the location of the dA→7 mutations, these results are consistent with the N1–N7 base pairing of the transWH pairs. Together, the simulation and base substitution experiments predict that the coralyne-induced homo-(dA) duplex structure adopts the transWH geometry.  相似文献   

5.
We have determined the crystal structure of the RNA octamer duplex r(guguuuac)/r(guaggcac) with a tandem wobble pair, G·G/U·U (motif III), to compare it with U·G/G·U (motif I) and G·U/U·G (motif II) and to better understand their relative stabilities. The crystal belongs to the rhombohedral space group R3. The hexagonal unit cell dimensions are a = b = 41.92 Å, c = 56.41 Å, and γ = 120°, with one duplex in the asymmetric unit. The structure was solved by the molecular replacement method at 1.9 Å resolution and refined to a final R factor of 19.9% and Rfree of 23.3% for 2862 reflections in the resolution range 10.0–1.9 Å with F ≥ 2σ(F). The final model contains 335 atoms for the RNA duplex and 30 water molecules. The A-RNA stacks in the familiar head-to-tail fashion forming a pseudo-continuous helix. The uridine bases of the tandem U·G pairs have slipped towards the minor groove relative to the guanine bases and the uridine O2 atoms form bifurcated hydrogen bonds with the N1 and N2 of guanines. The N2 of guanine and O2 of uridine do not bridge the ‘locked’ water molecule in the minor groove, as in motifs I and II, but are bridged by water molecules in the major groove. A comparison of base stacking stabilities of motif III with motifs I and II confirms the result of thermodynamic studies, motif I > motif III > motif II.  相似文献   

6.
The proteins coded by bacteriophate T4 replication genes 32, 41, 43, 44, 45, 61, and 62 together can replicate phi X174 DNA templates very efficiently. The fidelity of this in vitro replication reaction has been measured using an infectivity assay. The product molecules have the same specific infectivity as the template DNA. When an amber mutant DNA template is used, no increase in the frequency of revertants is seen even after more than 60 duplications in vitro. By using imbalances in the concentrations of deoxynucleotide substrates, the error rate during DNA replication in vitro can be greatly increased. Control experiments indicate that the increased mutagenesis is not due to the presence of dITP or dUTP as contaminants in the deoxynucleotide substrates used. The increase in the frequency of revertants is linearly related to the ratio of the correct and the incorrect deoxynucleotides. Determination of the DNA sequence of the revertants induced shows that a change in DNA sequence of the amber site predicted from the nucleotide bias occurs. DNA synthesis in vitro resembles in vivo replication in that the error rate depends not only upon the base change required for reversion but also upon the neighboring DNA sequences. The error rate is estimated to be 5 X 10(-6) at am3 site, 6.4 X 10(-7) at am86 site, and less than 2.9 X 10(-7) at am9 site. Comparison of the frequency of G-T and A-C mispairs reveals that most AT leads to GC transition mutations occur through G-T mispairs. Measurement of the frequency of the mispairs required to induce transversion mutations reveals that these occur primarily through purine-purine mispairs. Transition mutations are more frequent than transversion mutations at both the am3 and the am86 sites. These observations support the models for base pairing errors proposed by Watson and Crick ((1953) Nature 171, 964-967) and Topal and Fresco ((1976) Nature 263, 285-289).  相似文献   

7.
8.
RsrI [N6-adenine] DNA methyltransferase (M·RsrI), which recognizes GAATTC and is a member of a restriction–modification system in Rhodobacter sphaeroides, was purified to >95% homogeneity using a simplified procedure involving two ion exchange chromatographic steps. Electrophoretic gel retardation assays with purified M·RsrI were performed on unmethylated, hemimethylated, dimethylated or non-specific target DNA duplexes (25 bp) in the presence of sinefungin, a potent inhibitory analog of AdoMet. M·RsrI binding was affected by the methylation status of the DNA substrate and was enhanced by the presence of the cofactor analog. M·RsrI bound DNA substrates in the presence of sinefungin with decreasing affinities: hemimethylated > unmethylated > dimethylated >> non-specific DNA. Gel retardation studies with DNA substrates containing an abasic site substituted for the target adenine DNA provided evidence consistent with M·RsrI extruding the target base from the duplex. Consistent with such base flipping, an ~1.7-fold fluorescence intensity increase was observed upon stoichiometric addition of M·RsrI to hemimethylated DNA containing the fluorescent analog 2-aminopurine in place of the target adenine. Pre-steady-state kinetic and isotope- partitioning experiments revealed that the enzyme displays burst kinetics, confirmed the catalytic competence of the M·RsrI–AdoMet complex and eliminated the possibility of an ordered mechanism where DNA is required to bind first. The equilibrium dissociation constants for AdoMet, AdoHcy and sinefungin were determined using an intrinsic tryptophan fluorescence-quenching assay.  相似文献   

9.
The inhibitory effect of the polypeptide antibiotics netropsin and distamycin A on DNA dependent nucleic acid synthesis has been shown to be related to the base composition of the template DNA. A number of natural DNA's of quite different dA·dT content as well as poly (dI-dC)·poly (dI-dC), poly (dA-dT)·poly (dA-dT), poly (dA) · poly (dT) and poly (dG)·poly(dC) has been studied as templates in DNA and in part in RNA polymerase reaction. The highest binding efficiency of netropsin existing for (dA·dT)-containing DNA polymers and the less pronounced interaction with the (dI·dC)-containing polymer shown by the melting and CD spectral behaviour of the complexes are entirely reflected in the template inactivation. The same is evident for distamycin A. However, in contrast to netropsin the antibiotic distamycin A exhibits some binding tendency to poly (dG)·poly (dC). Binding effects of a netropsin derivative to DNA and (dA·dT)-containing polymers suggest the importance of hydrogen bonds of the peptide groups in the complex formation.  相似文献   

10.
1. Two species of double-helical RNA isolated from mycelium of Penicillium chrysogenum were titrated with acid at 25°C and 95°C (solvent 0.1m-sodium phosphate buffer). At 25°C denaturation occurred at about pH3. At 95°C in the denatured form cytosine residues titrated as a simple monobasic acid of pK3.9 compared with pK2.5 for the native form at 25°C. 2. On thermal denaturation in neutral and acidic solutions one species of RNA (38% rG·rC) `melted' in three distinct stages, equivalent to a mixture of three species, namely one of about 25% rG·rC, another of about 33% rG·rC and a third of about 46% rG·rC: the relative proportions were 0.25:0.35:0.40. 3. On thermal denaturation in acidic solutions the increase in the fraction of ionized cytosine residues concomitant with the `melting' of rG·rC base pair also affects the spectrum especially at 280nm and serves to enhance the contribution of rG·rC base pairs at this wavelength. The increment in ε(P) at 280nm on `melting' an rG·rC base pair approaches 53501·mol−1·cm−1 depending on pH, compared with 33501·mol−1·cm−1 at pH7. In contrast ε(P) at 280nm is scarcely affected by `melting' rA·rU base pairs or by the protonization of adenine residues. 4. Changes in the spectrum of Escherichia coli rRNA on denaturation in acidic solutions were studied to yield the mole fractions of rA·rU and rG·rC base pairs `melting' at particular pH values.  相似文献   

11.
Thymine DNA Glycosylase (TDG) performs essential functions in maintaining genetic integrity and epigenetic regulation. Initiating base excision repair, TDG removes thymine from mutagenic G·T mispairs caused by 5-methylcytosine (mC) deamination and other lesions including uracil (U) and 5-hydroxymethyluracil (hmU). In DNA demethylation, TDG excises 5-formylcytosine (fC) and 5-carboxylcytosine (caC), which are generated from mC by Tet (ten–eleven translocation) enzymes. Using improved crystallization conditions, we solved high-resolution (up to 1.45 Å) structures of TDG enzyme–product complexes generated from substrates including G·U, G·T, G·hmU, G·fC and G·caC. The structures reveal many new features, including key water-mediated enzyme–substrate interactions. Together with nuclear magnetic resonance experiments, the structures demonstrate that TDG releases the excised base from its tight product complex with abasic DNA, contrary to previous reports. Moreover, DNA-free TDG exhibits no significant binding to free nucleobases (U, T, hmU), indicating a Kd >> 10 mM. The structures reveal a solvent-filled channel to the active site, which might facilitate dissociation of the excised base and enable caC excision, which involves solvent-mediated acid catalysis. Dissociation of the excised base allows TDG to bind the beta rather than the alpha anomer of the abasic sugar, which might stabilize the enzyme–product complex.  相似文献   

12.
Pronounced instability of tandem IU base pairs in RNA   总被引:1,自引:1,他引:0       下载免费PDF全文
Optical melting was used to determine the stabilities of three series of RNA oligomers containing tandem XU base pairs, GGCXUGCC (5′XU3′), GGCUXGCC (5′UX3′) and GGCXXGGC/CCGUUCCG (5′XX3′), where X is either A, G or I (inosine). The helices containing tandem AU base pairs were the most stable in the first two series (5′XU3′ and 5′UX3′), with an average melting temperature ~11°C higher than the helices with tandem 5′GU3′ base pairs and 25°C higher than the helices with tandem 5′IU3′ base pairs. For the third series (5′XX3′), the helix containing tandem GG is the most stable, with an average melting temperature ~2°C higher than the helix with tandem AA base pairs and ~24°C higher than the helix with tandem II base pairs. The thermodynamic stability of the oligomers with tandem IU base pairs was also investigated as a function of magnesium ion concentration. As with normal A–U or G–U tandem duplexes, the data could best be interpreted as non-specific binding of magnesium ions to the inosine-containing RNA oligonucleotides.  相似文献   

13.
In previous papers of this series the temperature-dependent Raman spectra of poly(dA)·poly(dT) and poly(dA–dT)·poly(dA–dT) were used to characterize structurally the melting and premelting transitions in DNAs containing consecutive A·T and alternating A·T/T·A base pairs. Here, we describe procedures for obtaining thermodynamic parameters from the Raman data. The method exploits base-specific and backbone-specific Raman markers to determine separate thermodynamic contributions of A, T and deoxyribosyl-phosphate moieties to premelting and melting transitions. Key findings include the following: (i) Both poly(dA)·poly(dT) and poly(dA–dT)· poly(dA–dT) exhibit robust premelting transitions, due predominantly to backbone conformational changes. (ii) The significant van’t Hoff premelting enthalpies of poly(dA)·poly(dT) [ΔHvHpm = 18.0 ± 1.6 kcal·mol–1 (kilocalories per mole cooperative unit)] and poly(dA–dT)·poly(dA–dT) (ΔHvHpm = 13.4 ± 2.5 kcal·mol–1) differ by an amount (~4.6 kcal·mol–1) estimated as the contribution from three-centered inter-base hydrogen bonding in (dA)n·(dT)n tracts. (iii) The overall stacking free energy of poly(dA)· poly(dT) [–6.88 kcal·molbp–1 (kilocalories per mole base pair)] is greater than that of poly(dA–dT)· poly(dA–dT) (–6.31 kcal·molbp–1). (iv) The difference between stacking free energies of A and T is significant in poly(dA)·poly(dT) (ΔΔGst = 0.8 ± 0.3 kcal· molbp–1), but marginal in poly(dA–dT)·poly(dA–dT) (ΔΔGst = 0.3 ± 0.3 kcal·molbp–1). (v) In poly(dA)· poly(dT), the van’t Hoff parameters for melting of A (ΔHvHA = 407 ± 23 kcal·mol–1, ΔSvHA = 1166 ± 67 cal·°K–1·mol–1, ΔGvH(25°C)A = 60.0 ± 3.2 kcal·mol–1) are clearly distinguished from those of T (ΔHvHT = 185 ± 38 kcal·mol–1, ΔSvHT = 516 ± 109 cal·°K–1·mol–1, ΔGvH(25°C)T = 27.1 ± 5.5 kcal·mol–1). (vi) Similar relative differences are observed in poly(dA–dT)· poly(dA–dT) (ΔHvHA = 333 ± 54 kcal·mol–1, ΔSvHA = 961 ± 157 cal·°K–1·mol–1, ΔGvH(25°C)A = 45.0 ± 7.6 kcal· mol–1; ΔHvHT = 213 ± 30 kcal·mol–1, ΔSvHT = 617 ± 86 cal·°K–1·mol–1, ΔGvH(25°C)T = 29.3 ± 4.9 kcal·mol–1). The methodology employed here distinguishes thermodynamic contributions of base stacking, base pairing and backbone conformational ordering in the molecular mechanism of double-helical B DNA formation.  相似文献   

14.
Several cellular processes involve alignment of three nucleic acids strands, in which the third strand (DNA or RNA) is identical and in a parallel orientation to one of the DNA duplex strands. Earlier, using 2-aminopurine as a fluorescent reporter base, we demonstrated that a self-folding oligonucleotide forms a recombination-like structure consistent with the R-triplex. Here, we extended this approach, placing the reporter 2-aminopurine either in the 5′- or 3′-strand. We obtained direct evidence that the 3′-strand forms a stable duplex with the complementary central strand, while the 5′-strand participates in non-Watson–Crick interactions. Substituting 2,6-diaminopurine or 7-deazaadenine for adenine, we tested and confirmed the proposed hydrogen bonding scheme of the A*(T·A) R-type triplet. The adenine substitutions expected to provide additional H-bonds led to triplex structures with increased stability, whereas the substitutions consistent with a decrease in the number of H-bonds destabilized the triplex. The triplex formation enthalpies and free energies exhibited linear dependences on the number of H-bonds predicted from the A*(T·A) triplet scheme. The enthalpy of the 10 nt long intramolecular triplex of −100 kJ·mol−1 demonstrates that the R-triplex is relatively unstable and thus an ideal candidate for a transient intermediate in homologous recombination, t-loop formation at the mammalian telomere ends, and short RNA invasion into a duplex. On the other hand, the impact of a single H-bond, 18 kJ·mol−1, is high compared with the overall triplex formation enthalpy. The observed energy advantage of a ‘correct’ base in the third strand opposite the Watson–Crick base pair may be a powerful mechanism for securing selectivity of recognition between the single strand and the duplex.  相似文献   

15.
The interaction of the phage T4 Dam DNA-[N6-adenine] methyltransferase with 24mer synthetic oligonucleotide duplexes having different purine base substitutions in the palindromic recognition sequence, GATC, was investigated by means of gel shift and methyl transfer assays. The substitutions were introduced in either the upper or lower strand: guanine by 7-deazaguanine (G-->D) or 2-aminopurine (G-->N) and target adenine by purine (A-->P) or 2-aminopurine (A-->N). The effects of each base modification on binding/methylation were approximately equivalent for both strands. G-->D and G-->N substitutions resulted in a sharp decrease in binary complex formation. This suggests that T4 Dam makes hydrogen bonds with either the N7- or O6-keto groups (or both) in forming the complex. In contrast, A-->P and A-->N substitutions were much more tolerant for complex formation. This confirms our earlier observations that the presence of intact 5'-G:C base pairs at both ends of the methylation site is critical, but that base substitutions within the central A:T base pairs show less inhibition of complex formation. Addition of T4 Dam to a complete substrate mixture resulted in a burst of [3H]methylated product. In all cases the substrate dependencies of bursts and methylation rates were proportional to each other. For the perfect 24mer k cat = 0.014/s and K m = 7.7 nM was obtained. In contrast to binary complex formation the two guanine substitutions exerted relatively minor effects on catalytic turnover (the k cat was reduced at most 2. 5-fold), while the two adenine substitutions showed stronger effects (5- to 15-fold reduction in k cat). The effects of base analog substitutions on K m(DNA) were more variable: A-->P (decreased); A-->N and G-->D (unchanged); G-->N (increased).  相似文献   

16.
The alpha-sarcin loop of large subunit rRNAs is one of the sites of interaction of elongation factors with the ribosome, and the target of the cytotoxins alpha-sarcin and ricin. Using a genetic selection for increased frameshifting in a reporter gene, we have isolated a C --> U mutation at position 2666 in the alpha-sarcin loop. In the NMR-derived structure of the loop, bases equivalent to 2666 and 2654 are paired via a non-canonical base pairing interaction. Each of the three base substitutions at C2666 and A2654 was constructed by site-directed mutagenesis of a plasmid borne copy of the rrnB operon of Escherichia coli. Only the C2666 --> U and A2654 --> G mutations that resulted in the formation of canonical A-U and C-G base pairs respectively, increased the levels of stop codon readthrough and frameshifting. The effects of different base pair combinations at positions 2666 and 2654 on ribosome function were then tested by constructing and analyzing all possible base combinations at these sites. All A --> G base substitution mutations at position 2654 and C --> U substitutions at position 2666 increased the levels of translational errors. However, these effects were greatest when G2654 and U2666 had the potential to engage in standard Watson-Crick base pairing interactions. These data indicate that base identity as well as base pairing interactions are important for the function of this essential component of the large subunit rRNA.  相似文献   

17.
Mutations in the CFTR gene that lead to premature stop codons or splicing defects cause cystic fibrosis (CF) and are not amenable to treatment by small-molecule modulators. Here, we investigate the use of adenine base editor (ABE) ribonucleoproteins (RNPs) that convert A•T to G•C base pairs as a therapeutic strategy for three CF-causing mutations. Using ABE RNPs, we corrected in human airway epithelial cells premature stop codon mutations (R553X and W1282X) and a splice-site mutation (3849 + 10 kb C > T). Following ABE delivery, DNA sequencing revealed correction of these pathogenic mutations at efficiencies that reached 38–82% with minimal bystander edits or indels. This range of editing was sufficient to attain functional correction of CFTR-dependent anion channel activity in primary epithelial cells from CF patients and in a CF patient-derived cell line. These results demonstrate the utility of base editor RNPs to repair CFTR mutations that are not currently treatable with approved therapeutics.  相似文献   

18.
The HIV-1 type dimerization initiation signal (DIS) loop was used as a starting point for the analysis of the stability of Watson–Crick (WC) base pairs in a tertiary structure context. We used ultraviolet melting to determine thermodynamic parameters for loop–loop tertiary interactions and compared them with regular secondary structure RNA helices of the same sequences. In 1 M Na+ the loop–loop interaction of a HIV-1 DIS type pairing is 4 kcal/mol more stable than its sequence in an equivalent regular and isolated RNA helix. This difference is constant and sequence independent, suggesting that the rules governing the stability of WC base pairs in the secondary structure context are also valid for WC base pairs in the tertiary structure context. Moreover, the effect of ion concentration on the stability of loop–loop tertiary interactions differs considerably from that of regular RNA helices. The stabilization by Na+ and Mg2+ is significantly greater if the base pairing occurs within the context of a loop–loop interaction. The dependence of the structural stability on salt concentration was defined via the slope of a Tm/log [ion] plot. The short base-paired helices are stabilized by 8°C/log [Mg2+] or 11°C/log [Na+], whereas base-paired helices forming tertiary loop–loop interactions are stabilized by 16°C/log [Mg2+] and 26°C/log [Na+]. The different dependence on ionic strength that is observed might reflect the contribution of specific divalent ion binding to the preformation of the hairpin loops poised for the tertiary kissing loop–loop contacts.  相似文献   

19.
The thermodynamic contributions of rA·dA, rC·dC, rG·dG and rU·dT single internal mismatches were measured for 54 RNA/DNA duplexes in a 1 M NaCl buffer using UV absorbance thermal denaturation. Thermodynamic parameters were obtained by fitting absorbance versus temperature profiles using the curve-fitting program Meltwin. The weighted average thermodynamic data were fit using singular value decomposition to determine the eight non-unique nearest-neighbor parameters for each internal mismatch. The new parameters predict the ΔG°37, ΔH° and melting temperature (Tm) of duplexes containing these single mismatches within an average of 0.33 kcal/mol, 4.5 kcal/mol and 1.4°C, respectively. The general trend in decreasing stability for the single internal mismatches is rG·dG > rU·dT > rA·dA > rC·dC. The stability trend for the base pairs 5′ of the single internal mismatch is rG·dC > rC·dG > rA·dT > rU·dA. The stability trend for the base pairs 3′ of the single internal mismatch is rC·dG > rG·dC >> rA·dT > rU·dA. These nearest-neighbor values are now a part of a complete set of single internal mismatch thermodynamic parameters for RNA/DNA duplexes that are incorporated into the nucleic acid assay development software programs Visual oligonucleotide modeling platform (OMP) and ThermoBLAST.  相似文献   

20.
Triplex-forming oligodeoxynucleotide 15mers, designed to bind in the antiparallel triple-helical binding motif, containing single substitutions (Z) of the four isomeric αN7-, βN7-, αN9- and βN9-2-aminopurine (ap)-deoxyribonucleosides were prepared. Their association with double-stranded DNA targets containing all four natural base pairs (X-Y) opposite the aminopurine residues was determined by quantitative DNase I footprint titration in the absence of monovalent metal cations. The corresponding association constants were found to be in a rather narrow range between 1.0 × 106 and 1.3 × 108 M–1. The following relative order in Z × X-Y base-triple stabilities was found: Z = αN7ap: T-A > A-T> C-G ~ G-C; Z = βN7ap: A-T > C-G > G-C > T-A; Z = αN9ap: A-T = G-C > T-A > C-G; and Z = βN9ap: G-C > A-T > C-G > T-A.  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