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.
High precision ultrasonic and densimetric techniques have been used to study the interaction of Ni2+ ions with right-handed poly[d(G-C)]·poly[d(G-C)], poly-[d(A-C)]·poly[d(G-T)] and poly[d(A-T)]·poly[d(A-T)] in 5 mM CsCl, 0.2 mM HEPES, pH 7.5 at 20°C. From these measurements the changes in the apparent molar volume and the apparent molar adiabatic compressibility due to the interaction have been obtained. The volume effects of the binding, calculated per mole of Ni2+ ions, range from 11.7 to 23.9 cm3 mol–1 and the compressibility effects range from 19.3 × 10–4 to 43.1 × 10–4 cm3 mol–1 bar–1. These data are interpreted in terms of dehydration of the polynucleotides and Ni2+ ions, i.e. the release of water molecules from the hydration shells of the molecules. An increase in G+C content gives an increase in volume and compressibility effects, indicating a rise in the extent of dehydration. The dehydration effects of Ni2+ binding to poly[d(G-C)]·poly[d(G-C)] are approximately twice those of poly[d(A-T)]·poly[d(A-T)]. The volume and compressibility effects of Ni2+–EDTA complex formation have also been measured and used as a model system for quantitative estimation. These values revealed that Ni2+ ions can coordinate two atomic groups of poly[d(G-C)]·poly[d(G-C)], while in the case of the Ni2+–poly[d(A-T)]·poly[d(A-T)] complex volume and compressibility effects correspond to one direct or two indirect (through water) contacts.  相似文献   

2.
Wu J  Zhu BB  Yu J  Zhu H  Qiu L  Kindy MS  Gu L  Seidel A  Li GM 《Nucleic acids research》2003,31(22):6428-6434
Benzo[c]phenanthrene dihydrodiol epoxide (B[c] PhDE) is well known as an important environmental chemical carcinogen that preferentially modifies DNA in adenine residues. However, the molecular mechanism by which B[c]PhDE induces tumorigenesis is not fully understood. In this report, we demonstrate that DNA mismatch repair (MMR), a genome maintenance system, plays an important role in B[c]PhDE-induced carcinogensis by promoting apoptosis in cells treated with B[c]PhDE. We show that purified human MMR recognition proteins, MutSα and MutSβ, specifically recognized B[c]PhDE-DNA adducts. Cell lines proficient in MMR exhibited several-fold more sensitivity to killing than cell lines defective in either MutSα or MutLα by B[c]PhDE; the nature of this sensitivity was shown to be due to increased apoptosis. Additionally, wild-type mice exposed to B[c]PhDE had intestinal crypt cells that underwent apoptosis significantly more often than intestinal crypt cells found in B[c]PhDE-treated Msh2–/– or Mlh1–/– mice. These findings, combined with previous studies, suggest that the MMR system may serve as a general sensor for chemical-caused DNA damage to prevent damaged cells from mutagenesis and carcinogenesis by promoting apoptosis.  相似文献   

3.
A key issue in the nucleotide excision repair (NER) of bulky carcinogen–DNA adducts is the ability of the NER machinery to recognize and repair certain adducts while failing to repair others. Unrepaired adducts can survive to cause mutations that initiate the carcinogenic process. Benzo[c]phenanthrene (B[c]Ph), a representative fjord region polycyclic aromatic hydrocarbon, can be metabolically activated to the enantiomeric benzo[c]phenanthrene diol epoxides (B[c]PhDEs), (+)-(1S,2R,3R,4S)-3,4- dihydroxy-1,2-epoxy-1,2,3,4-tetrahydrobenzo[c]phe nanthrene and the corresponding (–)-(1R,2S,3S,4R) isomer. These react predominantly with adenine residues in DNA to produce the stereoisomeric 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts. Duplexes containing the 1R (+) or 1S (–) B[c]Ph-dA adduct in codon 61 of the human N-ras mutational hotspot sequence CA*A, with B[c]Ph modification at A*, are not repaired by the human NER system. However, the analogous stereoisomeric DNA adducts of the bay region benzo[a]pyrene diol epoxide (B[a]PDE), 10S (+)- and 10R (–)-trans-anti-B[a]P-N6-dA, are repaired in the same base sequence. In order to elucidate structural and thermodynamic origins of this phenomenon, we have carried out a 2 ns molecular dynamics simulation for the 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts in an 11mer duplex containing the human N-ras codon 61 sequence, and compared these results with our previous study of the B[a]P-dA adducts in the same sequence. The molecular mechanics Poisson– Boltzmann surface area (MM-PBSA) method was applied to calculate the free energies of the pair of stereoisomeric B[c]Ph-dA adducts, and a detailed structural analysis was carried out. The different repair susceptibilities of the B[a]P-dA adducts and the B[c]Ph-dA adducts can be attributed to different degrees of distortion, stemming from combined effects of differences in the quality of Watson–Crick hydrogen bonding, unwinding, stretching and helix backbone perturbations. These differences are due to the different intrinsic topologies of the rigid, planar bay region adducts versus the twisted, sterically hindered fjord region adducts.  相似文献   

4.
The interaction of hexamminecobalt(III), Co(NH3)63+, with 160 and 3000–8000 bp length calf thymus DNA has been investigated by circular dichroism, acoustic and densimetric techniques. The acoustic titration curves of 160 bp DNA revealed three stages of interaction: (i) Co(NH3)63+ binding up to the molar ratio [Co(NH3)63+]/[P] = 0.25, prior to DNA condensation; (ii) a condensation process between [Co(NH3)63+]/[P] = 0.25 and 0.30; and (iii) precipitation after [Co(NH3)63+]/[P] = 0.3. In the case of 3000–8000 bp DNA only two processes were observed: (i) binding up to [Co(NH3)63+]/[P] = 0.3; and (ii) precipitation after this point. In agreement with earlier observations, long DNA aggregates without changes in its B-form circular dichroism spectrum, while short DNA demonstrates a positive B→Ψ transition after [Co(NH3)63+]/[P] = 0.25. From ultrasonic and densimetric measurements the effects of Co(NH3)63+ binding on volume and compressibility have been obtained. The binding of Co(NH3)63+ to both short and long DNA is characterized by similar changes in volume and compressibility calculated per mole Co(NH3)63+: ΔV = 9 cm3 mol–1 and Δκ = 33 × 10–4 cm3 mol–1 bar–1. The positive sign of the parameters indicates dehydration, i.e. water release from Co(NH3)63+ and the atomic groups of DNA. This extent of water displacement would be consistent with the formation of two direct, hydrogen bonded contacts between the cation and the phosphates of DNA.  相似文献   

5.
Several studies have shown improved efficacy of cholesteryl-conjugated phosphorothioate antisense oligodeoxynucleotides. To gain insight into the mechanisms of the improved efficacy in vivo, we investigated the disposition of ISIS-9388, the 3′-cholesterol analog of the ICAM-1-specific phosphoro­thioate oligodeoxynucleotide ISIS-3082, in rats. Intravenously injected [3H]ISIS-9388 was cleared from the circulation with a half-life of 49.9 ± 2.2 min (ISIS-3082, 23.3 ± 3.8 min). At 3 h after injection, the liver contained 63.7 ± 3.3% of the dose. Compared to ISIS-3082, the hepatic uptake of ISIS-9388 is ~2-fold higher. Endothelial, Kupffer and parenchymal cells accounted for 45.7 ± 5.7, 33.0 ± 5.9 and 21.3 ± 2.6% of the liver uptake of [3H]ISIS-9388, respectively, and intracellular concentrations of ~2, 75 and 50 µM, respectively, could be reached in these cells (1 mg/kg dose). Preinjection with polyinosinic acid or poly­adenylic acid reduced the hepatic uptake of [3H]ISIS-9388, which suggests the involvement of (multiple) scavenger receptors. Size exclusion chromatography of mixtures of the oligonucleotides and rat plasma indicated that ISIS-9388 binds to a larger extent to high molecular weight proteins than ISIS-3082. Analysis by agarose gel electrophoresis indicated that ISIS-9388 binds more tightly to plasma proteins than ISIS-3082. The different interaction of the oligonucleotides with plasma proteins possibly explains their different dispositions. We conclude that cholesterol conjugation results in high accumulation of phosphorothioate oligodeoxynucleotides in various liver cell types, which is likely to be beneficial for antisense therapy of liver-associated diseases.  相似文献   

6.
Eckert KA  Yan G 《Nucleic acids research》2000,28(14):2831-2838
Mutagenesis at [GT/CA]10, [TC/AG]11 and [TTCC/AAGG]9 microsatellite sequences inserted in the herpes simplex virus thymidine kinase (HSV-tk) gene was analyzed in isogenic mutL+ and mutL Escherichia coli. In both strains, significantly more expansion than deletion mutations were observed at the [TTCC/AAGG]9 motif relative to either dinucleo­tide motif. As the HSV-tk coding sequence contains an endogenous [G/C]7 mononucleotide repeat and ~1000 bp of unique sequence, we were able to compare mutagenesis among various sequence motifs. We observed that the relative risk of mutation in E.coli is: [TTCC/AAGG]9 > [GT/CA]10 ~ [TC/AG]11 > unique ~ [G/C]7. The mutation frequency varied 1400-fold in mutL+ cells between the tetranucleotide motif and the mononucleotide motif, but only 50-fold in mutL cells. The [G/C]7 sequence was destabilized the greatest and the tetranucleotide motif the least by loss of mismatch repair. These results demonstrate that the quantitative risk of mutation at various microsatellites greatly depends on the DNA sequence composition. We suggest alternative models for the production of expansion mutations during lagging strand replication of the [TTCC/AAGG]9 microsatellite.  相似文献   

7.
Magnesium ions strongly influence the structure and biochemical activity of RNA. The interaction of Mg2+ with an equimolar mixture of poly(rA) and poly(rU) has been investigated by UV spectroscopy, isothermal titration calorimetry, ultrasound velocimetry and densimetry. Measurements in dilute aqueous solutions at 20°C revealed two differ ent processes: (i) Mg2+ binding to unfolded poly(rA)·poly(rU) up to [Mg2+]/[phosphate] = 0.25; and (ii) poly(rA)·2poly(rU) triplex formation at [Mg2+]/[phosphate] between 0.25 and 0.5. The enthalpies of these two different processes are favorable and similar to each other, ~–1.6 kcal mol–1 of base pairs. Volume and compressibility effects of the first process are positive, 8 cm3 mol–1 and 24 × 10–4 cm3 mol–1 bar–1, respectively, and correspond to the release of water molecules from the hydration shells of Mg2+ and the polynucleotides. The triplex formation is also accompanied by a positive change in compressibility, 14 × 10–4 cm3 mol–1 bar–1, but only a small change in volume, 1 cm3 mol–1. A phase diagram has been constructed from the melting experiments of poly(rA)·poly(rU) at a constant K+ concentration, 140 mM, and various amounts of Mg2+. Three discrete regions were observed, corresponding to single-, double- and triple-stranded complexes. The phase boundary corresponding to the transition between double and triple helical conformations lies near physiological salt concentrations and temperature.  相似文献   

8.
The cytosine–cytosine (C–C) pair is one of the least stable DNA mismatch pairs. The bases of the C–C mismatch are only weakly hydrogen bonded, and previous work has shown that, in certain sequence contexts, they can become unstacked from the core helix, and adopt an ‘extrahelical’ location. Here, using DNA duplexes with d[GCC]n·d[GCC]n fragments containing C–C mismatches in a 1,4 bp relationship, we show that cytosine bases of different formal mismatch pairs can be crosslinked by mechlorethamine. For example, in the duplex d[CTCTCGCCGCCGCCGTATC]·d[GATACGCCGCCGCCGAGAG], where underlined cytosine bases are present as the formal C–C mismatch pairs C7–C32, C10–C29 and C13–C26, we show that two mechlorethamine crosslinks form between C13 and C29 and between C10 and C32, in addition to crosslinks at C7–C32, C10–C29 and C13–C26 (we have reported previously the crosslinking of formal C–C pairs by mechlorethamine). We interpret the formation of the C13–C29 and C10–C32 crosslinks as evidence of an extrahelical location of the crosslinkable cytosines. Such extrahelical cytosine bases have been observed previously for a single C–C mismatch pair (in the so-called E-motif conformation). In the E-motif, the extrahelical cytosines are folded back towards the 5′-end of the duplex, consistent with our crosslinking data, and also consistent with the absence of C7–C29 and C10–C26 crosslinks in the current work. Hence, our data provide evidence for an extended E-motif DNA (eE-DNA) conformation in short d[GCC]n·d[GCC]n repeat fragments, and raise the possibility that such structures might occur in much longer d[GCC]n·d[GCC]n repeat tracts.  相似文献   

9.
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   

10.
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   

11.
The solution structure and hydration of the chimeric duplex [d(CGC)r(aaa)d(TTTGCG)]2, in which the central hybrid segment is flanked by DNA duplexes at both ends, was determined using two-dimensional NMR, simulated annealing and restrained molecular dynamics. The solution structure of this chimeric duplex differs from the previously determined X-ray structure of the analogous B-DNA duplex [d(CGCAAATTTGCG)]2 as well as NMR structure of the analogous A-RNA duplex [r(cgcaaauuugcg)]2. Long-lived water molecules with correlation time τc longer than 0.3 ns were found close to the RNA adenine H2 and H1′ protons in the hybrid segment. A possible long-lived water molecule was also detected close to the methyl group of 7T in the RNA–DNA junction but not with the other two thymines (8T and 9T). This result correlates with the structural studies that only DNA residue 7T in the RNA–DNA junction adopts an O4′-endo sugar conformation, while the other DNA residues including 3C in the DNA–RNA junction, adopt C1′-exo or C2′-endo conformations. The exchange rates for RNA C2′-OH were found to be ~520 s–1. This slow exchange rate may be due to the narrow minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2, which may trap the water molecules and restrict the dynamic motion of hydroxyl protons. The minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2 is wider than its B-DNA analog but narrower than that of the A-RNA analog. It was further confirmed by its titration with the minor groove binding drug distamycin. A possible 2:1 binding mode was found by the titration experiments, suggesting that this chimeric duplex contains a wider minor groove than its B-DNA analog but still narrow enough to hold two distamycin molecules. These distinct structural features and hydration patterns of this chimeric duplex provide a molecular basis for further understanding the structure and recognition of DNA·RNA hybrid and chimeric duplexes.  相似文献   

12.
Ensemble and single-molecule fluorescence measurements of 2′-(4-hydroxyphenyl)-5-[5-(4-methylpiperazine-1-yl) benzimidazo-2-yl]-benzimidazole (H-258)– calf thymus (CT) DNA complexes at various [H-258]/[DNA bp] ratios were performed to elucidate the binding of H-258 with DNA. Upon binding to double-stranded CT DNA (CT ds DNA) at a [H-258]/[DNA bp] ratio of 0.05 the relative fluorescence quantum yield, Φf, of H-258 increases from 0.02 to 0.58. The fluorescence decay can be fitted almost by a mono-exponential model with a lifetime of ~3.6 ns. This indicates that H-258 binds almost quantitatively in the minor groove of DNA at low [H-258]/[DNA bp] ratios. With increasing [H-258]/[DNA bp] ratios, e.g. 0.15 and 0.20, the fluorescence quantum yield of H-258 decreases to 0.28 and 0.19, respectively. Fitting of the fluorescence decays measured for higher [H-258]/[DNA bp] ratios reveals the presence of additional shorter fluorescence lifetime components in the range of 0.5–2.0 ns. Our results suggest that H-258 partially intercalates in G:C sequences at higher [H-258]/[DNA bp] ratios reflected by a lifetime component of 1.5–2 ns. In addition, stacking or adsorption of H-258 molecules on DNA occurs at higher [H-258]/[DNA bp] ratios. These molecules exhibit a short fluorescence lifetime of ~500 ps and are more exposed to the aqueous environment. Fluorescence transients of the intensity and lifetime of single H-258 CT ds DNA demonstrate that weakly (unspecific) bound H-258 molecules exhibit a shorter fluorescence lifetime and a strongly reduced photostability.  相似文献   

13.
Polymerization of fibrin, the primary structural protein of blood clots and thrombi, occurs through binding of knobs ‘A’ and ‘B’ in the central nodule of fibrin monomer to complementary holes ‘a’ and ‘b’ in the γ- and β-nodules, respectively, of another monomer. We characterized the A:a and B:b knob-hole interactions under varying solution conditions using molecular dynamics simulations of the structural models of fibrin(ogen) fragment D complexed with synthetic peptides GPRP (knob ‘A’ mimetic) and GHRP (knob ‘B’ mimetic). The strength of A:a and B:b knob-hole complexes was roughly equal, decreasing with pulling force; however, the dissociation kinetics were sensitive to variations in acidity (pH 5–7) and temperature (T = 25–37 °C). There were similar structural changes in holes ‘a’ and ‘b’ during forced dissociation of the knob-hole complexes: elongation of loop I, stretching of the interior region, and translocation of the moveable flap. The disruption of the knob-hole interactions was not an “all-or-none” transition as it occurred through distinct two-step or single step pathways with or without intermediate states. The knob-hole bonds were stronger, tighter, and more brittle at pH 7 than at pH 5. The B:b knob-hole bonds were weaker, looser, and more compliant than the A:a knob-hole bonds at pH 7 but stronger, tighter, and less compliant at pH 5. Surprisingly, the knob-hole bonds were stronger, not weaker, at elevated temperature (T = 37 °C) compared with T = 25 °C due to the helix-to-coil transition in loop I that helps stabilize the bonds. These results provide detailed qualitative and quantitative characteristics underlying the most significant non-covalent interactions involved in fibrin polymerization.  相似文献   

14.
Magnesium ions play important roles in the structure and function of nucleic acids. Whereas the tertiary folding of RNA often requires magnesium ions binding to tight places where phosphates are clustered, the molecular basis of the interactions of magnesium ions with RNA helical regions is less well understood. We have refined the crystal structures of four decamer oligonucleotides, d(ACCGGCCGGT), r(GCG)d(TATACGC), r(GC)d(GTATACGC) and r(G)d(GCGTATACGC) with bound hexahydrated magnesium ions at high resolution. The structures reveal that A-form nucleic acid has characteristic [Mg(H2O)6]2+ binding modes. One mode has the ion binding in the deep major groove of a GpN step at the O6/N7 sites of guanine bases via hydrogen bonds. Our crystallographic observations are consistent with the recent NMR observations that in solution [Co(NH3)6]3+, a model ion of [Mg(H2O)6]2+, binds in an identical manner. The other mode involves the binding of the ion to phosphates, bridging across the outer mouth of the narrow major groove. These [Mg(H2O)6]2+ ions are found at the most negative electrostatic potential regions of A-form duplexes. We propose that these two binding modes are important in the global charge neutralization, and therefore stability, of A-form duplexes.  相似文献   

15.
Methionine (Met) is an essential amino acid that is needed for the synthesis of S-adenosylmethionine (AdoMet), the major biological methylating agent. Methionine used for AdoMet synthesis can be replenished via remethylation of homocysteine. Alternatively, homocysteine can be converted to cysteine via the transsulfuration pathway. Aberrations in methionine metabolism are associated with a number of complex diseases, including cancer, anemia, and neurodegenerative diseases. The concentration of methionine in blood and in organs is tightly regulated. Liver plays a key role in buffering blood methionine levels, and an interesting feature of its metabolism is that parallel tracks exist for the synthesis and utilization of AdoMet. To elucidate the molecular mechanism that controls metabolic fluxes in liver methionine metabolism, we have studied the dependencies of AdoMet concentration and methionine consumption rate on methionine concentration in native murine hepatocytes at physiologically relevant concentrations (40–400 µM). We find that both [AdoMet] and methionine consumption rates do not change gradually with an increase in [Met] but rise sharply (~10-fold) in the narrow Met interval from 50 to 100 µM. Analysis of our experimental data using a mathematical model reveals that the sharp increase in [AdoMet] and the methionine consumption rate observed within the trigger zone are associated with metabolic switching from methionine conservation to disposal, regulated allosterically by switching between parallel pathways. This regulatory switch is triggered by [Met] and provides a mechanism for stabilization of methionine levels in blood over wide variations in dietary methionine intake.  相似文献   

16.
Poly(ADP-ribose) polymerase (PARP) knockout mice are resistant to murine models of human diseases such as cerebral and myocardial ischemia, traumatic brain injury, diabetes, Parkinsonism, endotoxic shock and arthritis, implicating PARP in the pathogenesis of these diseases. Potent selective PARP inhibitors are therefore being evaluated as novel therapeutic agents in the treatment of these diseases. Inhibition or depletion of PARP, however, increases genomic instability in cells exposed to genotoxic agents. We recently demonstrated the presence of a genomically unstable tetraploid population in PARP–/– fibroblasts and its loss after stable transfection with PARP cDNA. To elucidate whether the genomic instability is attributable to PARP deficiency or lack of PARP activity, we investigated the effects of PARP inhibition on development of tetraploidy. Immortalized wild-type and PARP–/– fibroblasts were exposed for 3 weeks to 20 µM GPI 6150 (1,11b-dihydro-[2H]benzopyrano[4,3,2-de]isoquinolin-3-one), a novel small molecule specific competitive inhibitor of PARP (Ki = 60 nM) and one of the most potent PARP inhibitors to date (IC50 = 0.15 µM). Although GPI 6150 initially decreased cell growth in wild-type cells, there was no effect on cell growth or viability after 24 h. GPI 6150 inhibited endogenous PARP activity in wild-type cells by ~91%, to about the residual levels in PARP–/– cells. Flow cytometric analysis of unsynchronized wild-type cells exposed for 3 weeks to GPI 6150 did not induce the development of tetraploidy, suggesting that, aside from its catalytic function, PARP may play other essential roles in the maintenance of genomic stability.  相似文献   

17.
Members of the ribonuclease III superfamily of double-strand-specific endoribonucleases participate in diverse RNA maturation and decay pathways. Ribonuclease III of the gram-negative bacterium Escherichia coli processes rRNA and mRNA precursors, and its catalytic action can regulate gene expression by controlling mRNA translation and stability. It has been proposed that E.coli RNase III can function in a non-catalytic manner, by binding RNA without cleaving phosphodiesters. However, there has been no direct evidence for this mode of action. We describe here an RNA, derived from the T7 phage R1.1 RNase III substrate, that is resistant to cleavage in vitro by E.coli RNase III but retains comparable binding affinity. R1.1[CL3B] RNA is recognized by RNase III in the same manner as R1.1 RNA, as revealed by the similar inhibitory effects of a specific mutation in both substrates. Structure-probing assays and Mfold analysis indicate that R1.1[CL3B] RNA possesses a bulge– helix–bulge motif in place of the R1.1 asymmetric internal loop. The presence of both bulges is required for uncoupling. The bulge–helix–bulge motif acts as a ‘catalytic’ antideterminant, which is distinct from recognition antideterminants, which inhibit RNase III binding.  相似文献   

18.
We report for the first time the in vitro characterization of a reverse tetracycline repressor (revTetR). The dimeric wild-type repressor (TetR) binds to tet operator tetO in the absence of the inducer anhydrotetracycline (atc) to confer tight repression. We have isolated the revTetR G96E L205S mutant, which, contrary to TetR, binds tetO only in the presence of atc. This reverse acting mutant was overproduced and purified. Effector and DNA binding properties were analyzed by EMSA and quantified by fluorescence titration and surface plasmon resonance. The association constant KA of revTetR for binding of [atcMg]+ is ~108 M–1, four orders of magnitude lower than that of TetR. The affinity of TetR for tetO is 5.6 ± 2 × 109 M–1 and that for revTetR in the presence of atc is 1 ± 0.2 × 108 M–1. Both induced forms, the atc-bound TetR and the free revTetR, have the same low affinity of 4 ± 1 × 105 M–1 for DNA. Therefore, atc does not act as a dimerization agent for revTetR. We discuss the structural differences between TetR and revTetR potentially underlying this reversal of activity.  相似文献   

19.
The cytosine base in DNA undergoes hydrolytic deamination at a considerable rate when UV radiation induces formation of a cyclobutane pyrimidine dimer (CPD) with an adjacent pyrimidine base. We have synthesized a phosphoramidite building block of a cissyn cyclobutane thymine–uracil dimer (T[]U), which is the deaminated form of the CPD at a TC site, and incorporated it into oligodeoxyribonucleotides. The previously reported method for synthesis of the thymine dimer (T[]T) was applied, using partially protected thymidylyl-(3′–5′)-2′-deoxyuridine as the starting material, and after triplet- sensitized irradiation, the configuration of the base moiety in the major product was determined by NMR spectroscopy. Presence of the cissyn cyclobutane dimer in the obtained oligonucleotides was confirmed by UV photoreversal and reaction with T4 endonuclease V. Using a 30mer containing T[]U, translesion synthesis by human DNA polymerase η was analyzed. There was no difference in the results between the templates containing T[]T and T[]U and pol η bypassed both lesions with the same efficiency, incorporating two adenylates. This enzyme showed fidelity to base pair formation, but this replication causes a C→T transition because the original sequence is TC.  相似文献   

20.
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   

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