首页 | 官方网站   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.
DNA bending induced by six DNA (cytosine-5) methyltransferases was studied using circular permutation gel mobility shift assay. The following bend angles were obtained: M.BspRI (GGm5CC), 46–50°; M.HaeIII (GGm5CC), 40–43°; M.SinI (GGWm5CC), 34–37°; M.Sau96I (GGNm5CC), 52–57°; M.HpaII (Cm5CGG), 30°; and M.HhaI (Gm5CGC), 13°. M.HaeIII was also tested with fragments carrying a methylated binding site, and it was found to induce a 32° bend. A phase-sensitive gel mobility shift assay, using a set of DNA fragments with a sequence-directed bend and a single methyltransferase binding site, indicated that M.HaeIII and M.BspRI bend DNA toward the minor groove. The DNA curvature induced by M.HaeIII contrasts with the lack of DNA bend observed for a covalent M.HaeIII–DNA complex in an earlier X-ray study. Our results and data from other laboratories show a correlation between the bending properties and the recognition specificities of (cytosine-5) methyltransferases: enzymes recognizing a cytosine 3′ to the target cytosine tend to induce greater bends than enzymes with guanine in this position. We suggest that the observed differences indicate different mechanisms employed by (cytosine-5) methyltransferases to stabilize the helix after the target base has flipped out.  相似文献   

2.
The gene coding for the SinI DNA-methyltransferase, a modification enzyme able to recognize and methylate the internal cytosine of the GGA/TCC sequence, was subjected to in vitro mutagenesis, DNA-shuffling and a strong selection for relaxed GGNCC recognition specificity. As a result of this in vitro evolution experiment, a mutant gene with the required phenotype was selected. The mutant SinI methyltransferase carried five amino acid substitutions. None of these was found in the ‘variable region’ that were thought to be responsible for sequence specificity. Three were located near the N-terminal end, preceding the first conserved structural motif of the enzyme; two were found between conserved motifs VI and VII. A clone engineered to carry out only the latter two replacements (L214S and Y229H) displays relaxed recognition specificity similar to that of the parental mutant, whereas the clone carrying only the N-terminal replacements showed a much weaker change in recognition specificity. The enzyme with two internal mutations was purified and characterized. Its catalytic activity (kcat/Km) was ~5-fold lower towards GGA/TCC and 20-fold higher towards GGG/CCC than that of the wild-type enzyme.  相似文献   

3.
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   

4.
Eco RII restriction endonuclease cleaves synthetic DNA-duplexes in which the recognition sites of this enzyme (5..CC T A GG...) are repeated every 9 base pairs with the alternating orientation of the central AT pair. It operates in a processive mode, i.e. the bound enzyme molecule slides along the substrate toward neighboring recognition sites. Nona-nucleotides are the main products of the cleavage. The data obtained point to the capability of Eco RII endonuclease to recognize and cleave the substrate under both possible orientations of the central AT-pair of the recognition site with respect to the bound enzyme molecule. These data also show the close similarity of DNA structures in a complex with theenzyme and without.  相似文献   

5.
Abstract

Oligonucleotides containing 1-(β-D-2′-deoxy-threo-pentofuranosyl)cytosine (dCx) and/or 1-(β-D-2′-deoxy-threo-pentofuranosyl)thymine (dTx) in place of dC and dT residues in the EcoRII and MvaI recognition site CCA/TGG were synthesized in order to investigate specific recognition of the DNA sugar-phosphate backbone by EcoRII and MvaI restriction endonucleases. In 2′-deoxyxylosyl moieties of dCx and dTx, 3′-hydroxyl groups were inverted, which perturbs the related individual phosphates. Introduction of a single 2′-deoxyxylo-syl moiety into a dC·dG pair resulted in a minor destabilization of double-stranded DNA structure. In the case of a dA·dT pair the effect of a 2′-deoxyxylose incorporation was much more pronounced. Multiple dCx modifications and their combination with dTx did not enhance the destabilization effect. Hydrolysis of dCx-containing DNA duplexes by EcoRII endonuclease was blocked and binding affinity was strongly depended on the location of an altered sugar. A DNA duplex containing a dTx residue was cleaved by the enzyme, but kcat/KM was slightly reduced. In contrast, MvaI endonuclease efficiently cleaved both types of sugar-altered substrate analogs. However it did not cleave conformationally perturbed scissile bonds, when the corresponding unmodified bonds were perfectly hydrolyzed in the same DNA duplexes. Based on these data the possible contributions of individual phosphates in the recognition site to substrate recognition and catalysis by EcoRII were proposed. We observed strikingly non-equivalent inputs for different phosphates with respect to their effect on EcoRII-DNA complex formation.  相似文献   

6.
A necessary feature of the natural base triads for triplex formation is the requirement of a purine (A or G) in the central position, since only these provide sets of two hydrogen bond donors/acceptors in the major groove of the double helix. Pyrimidine bases devoid of this feature have incompatible complementarity and lead to triplexes with lower stability. This paper demonstrates that 5-aminouracil (U#) (I), a pyrimidine nucleobase analogue of T in which 5-methyl is replaced by 5-amino group, with hydrogen bonding sites on both sides, is compatible in the central position of triplex triad X*U#·A, where X = A/G/C/T/2-aminopurine (AP), and * and · represent Hoogsteen and Watson–Crick hydrogen bonding patterns respectively. A novel recognition selectivity based on the orientation (parallel/antiparallel) of the third strand purines A, G or AP with A in the parallel motif (Ap*U#·A), and G/AP in the antiparallel motif (Gap/APap*U#·A) is observed. Similarly for pyrimidines in the third strand, C is accepted only in a parallel mode (Cp*U#·A). Significantly, T is recognised in both parallel and antiparallel modes (Tp/Tap*U#·A), with the antiparallel mode being stable compared to the parallel one. The ‘U#’ triplexes are also more stable than the corresponding control ‘T’ triplexes. The results expand the lexicon of triplex triads with a recognition motif consisting of pyrimidine in the central strand.  相似文献   

7.
The dynamics of the opening-closing of the constituent base-pairs as well as of the exchange kinetics of the base-paired imino and amino protons with water in a DNA-RNA hybrid, [5′r(G1A2U3U4G5A6A7)3′]:5′p[d(T8C9A10A11T12C13)]3′-Pzn] duplex (I), are reported here in details for the first time. The exchange kinetics of amino and imino protons in the DNA-RNA hybrid (duplex I) have been compared with identical studies on the following B-DNA duplexes: d(C1G2T3A4C5G6)2 (II), d[p(5′T1G2T3T4T5G6 G7C8)3′]:d[p(5′C9C10A11A12A13C14A15)3′] (III), d(C5G6C7G8A9A10T11T12C13G14C15G16)2 (IV) and d(C1G2C3G4C5G6C7G8A9A10T11T12C13G14C15G16C17G18C19G20)2 (V). This comparative study shows that the life-times τo of various base-pairs in the DNA-RNA hybrid (I) varies in the range of ∼ 1 ms, and they are quite comparable to those of the shorter B-DNA duplexes (II) and (III), but very different from the τo of the larger duplexes (IV) and (V): the τo for the base pair of T11 and T12 residues in the 20-mer (duplex V) are 2.9 ± 2.3 ms and 23.2 ± 8.9 ms, respectively, while the corresponding τo in the 12-mer (duplex IV) are 2.8 ± 2.2 ms and 17.4 ± 5.4 ms. It has also been shown that the total energy of activation (Ea) assessed from the exchange rates of both imino and amino protons, representing energetic contributions from both base-pair and helix opening-closing as well as from the exchange process of the imino protons from the open state with the bound water, is close to the Ea of the short B-DNA duplex (Ea ≈ 28–47 kcal/mol).  相似文献   

8.
ABSTRACT

Affinity modification of EcoRII DNA methyltransferase (M·EcoRII) by DNA duplexes containing oxidized 2′-O-β-D-ribofuranosylcytidine (Crib*) or 1-(β-D-galactopyranosyl)thymine (Tgal*) residues was performed. Cross-linking yields do not change irrespective of whether active Crib* replaces an outer or an inner (target) deoxycytidine within the EcoRII recognition site. Chemical hydrolysis of M·EcoRII in the covalent cross-linked complex with the Tgal*-substituted DNA indicates the region Gly268-Met391 of the methylase that is likely to interact with the DNA sugar-phosphate backbone. Both specific and non-specific DNA interact with the same M·EcoRII region. Our results support the theoretically predicted DNA binding region of M·EcoRII.  相似文献   

9.
The molecular mechanisms of DNA recognition and modification by EcoRII DNA methyltransferase (M.EcoRII) were studied using 14-mer substrate analogs containing 2-aminopurine or 1′,2′-dideoxy-D-ribofuranose in the M.EcoRII recognition site. The efficiency of DNA binding and methylation depended on the position of a modified nucleoside residue in the recognition site. A structural model of M.EcoRII in complex with substrate DNA and the cofactor analog S-adenosyl-L-homocysteine (AdoHcy) was constructed using the available crystal structures of M.Hha and M.HaeIII and the recent Frankenstein’s monster approach. The amino acid residues interacting with DNA were predicted based on the model. In addition, theoretical and experimental findings made it possible to predict the groups of atoms of the heterocyclic bases of the M.EcoRII recognition site that are presumably involved in the interactions with the enzyme.  相似文献   

10.
A Model for the Action of Vinblastine in Vivo   总被引:1,自引:0,他引:1       下载免费PDF全文
A model for the action of vinblastine (VLB) on cells multiplying exponentially in vivo with a generation time, TG, has been derived. It is based on the assumption that cells attempting to pass through mitosis in the presence of VLB lose their proliferative capacity and that this lethal effect occurs only when the cells are exposed to a concentration of VLB which is above a critical value, Ck. The model leads to two predictions. First, that the percentage of cells surviving at any time after exposure to a dose, D, of VLB is 100% if D < Dk and decreases to 0% after a time, TG, following a dose D ≥ Dk·2T G/T1/2, where Dk represents the dose of VLB required to produce the concentration Ck, and T1/2 is the half-life of the VLB in vivo. Second, that the time, TG, at which the percentage of cells surviving an exposure to VLB, at doses greater than Dk·2U G/T1/2, decreases to zero should be equal to the generation time of the cells. Both of these predictions were confirmed experimentally which indicates that the model adequately explains the action of VLB in vivo.  相似文献   

11.
The 2′-deoxynucleoside containing the synthetic base 1-[(2R,4S,5R)-4-hydroxy-5-(hydroxymethyl)-tetrahydrofuran-2-yl)-1H-perimidin-2(3H)-one] (dPer) recognizes in DNA the O6-benzyl-2′-deoxyguanosine nucleoside (O6-Bn-dG), formed by exposure to N-benzylmethylnitrosamine. Herein, we show how dPer distinguishes between O6-Bn-dG and dG in DNA. The structure of the modified Dickerson–Drew dodecamer (DDD) in which guanine at position G4 has been replaced by O6-Bn-dG and cytosine C9 has been replaced with dPer to form the modified O6-Bn-dG:dPer (DDD-XY) duplex [5′-d(C1G2C3X4A5A6T7T8Y9G10C11G12)-3′]2 (X = O6-Bn-dG, Y = dPer) reveals that dPer intercalates into the duplex and adopts the syn conformation about the glycosyl bond. This provides a binding pocket that allows the benzyl group of O6-Bn-dG to intercalate between Per and thymine of the 3′-neighbor A:T base pair. Nuclear magnetic resonance data suggest that a similar intercalative recognition mechanism applies in this sequence in solution. However, in solution, the benzyl ring of O6-Bn-dG undergoes rotation on the nuclear magnetic resonance time scale. In contrast, the structure of the modified DDD in which cytosine at position C9 is replaced with dPer to form the dG:dPer (DDD-GY) [5′-d(C1G2C3G4A5A6T7T8Y9G10C11G12)-3′]2 duplex (Y = dPer) reveals that dPer adopts the anti conformation about the glycosyl bond and forms a less stable wobble pairing interaction with guanine.  相似文献   

12.
The diversity of dinucleotide sequences at the 5′ ends of vaccinia virus mRNA's was determined by a two-dimensional electrophoresis procedure. RNA labeled with S-adenosyl[methyl-3H]methionine was synthesized in vitro by enzymes present in vaccinia virus cores. The RNA, ending in m7G(5′)pppNmpN−, was β-eliminated and treated with alkaline phosphatase. After digestion with RNases T2, T1, and A, all eight possible dinucleotides containing Gm and Am were identified. They are, in decreasing order of abundance: GmpUp (22%), AmpCp (18%), GmpAp (16%), GmpCp (15%), AmpAp (11%), AmpUp (10%), AmpGp (7%), and GmpGp (2%).  相似文献   

13.
A novel Gram-negative bacterium, designated ZWT, was isolated from a soil sample of the Western Desert of China, and its phenotypic properties and phylogenetic position were investigated using a polyphasic approach. Growth occurred on TGY medium at 5–42°C with an optimum of 30°C, and at pH 7.0–11.0 with an optimum of pH 9.0. The predominant cellular fatty acids were summed feature 3 (C16:1 ω7c/C16:1 ω6c or C16:1 ω6c/C16:1 ω7c) (39.22%), iso-C15:0 (27.91%), iso-C17:0 3OH (15.21%), C16:0 (4.98%), iso-C15:0 3OH (3.03%), C16:0 3OH (5.39%) and C14:0 (1.74%). The major polar lipid of strain ZWT is phosphatidylethanolamine. The only menaquinone observed was MK-7. The GC content of the DNA of strain ZWT is 44.9 mol%. rDNA phylogeny, genome relatedness and chemotaxonomic characteristics all indicate that strain ZWT represents a novel species of the genus Sphingobacterium. We propose the name S. deserti sp. nov., with ZWT (= KCTC 32092T = ACCC 05744T) as the type strain. Whole genome optical mapping and next-generation sequencing was used to derive a finished genome sequence for strain ZWT, consisting of a circular chromosome of 4,615,818 bp in size. The genome of strain ZWT features 3,391 protein-encoding and 48 tRNA-encoding genes. Comparison of the predicted proteome of ZWT with those of other sphingobacteria identified 925 species-unique proteins that may contribute to the adaptation of ZWT to its native, extremely arid and inhospitable environment. As the first finished genome sequence for any Sphingobacterium, our work will serve as a useful reference for subsequent sequencing and mapping efforts for additional strains and species within this genus.  相似文献   

14.
Single-stranded DNA or double-stranded DNA has the potential to adopt a wide variety of unusual duplex and hairpin motifs in the presence (trans) or absence (cis) of ligands. Several principles for the formation of those unusual structures have been established through the observation of a number of recurring structural motifs associated with different sequences. These include: (i) internal loops of consecutive mismatches can occur in a B-DNA duplex when sheared base pairs are adjacent to each other to confer extensive cross- and intra-strand base stacking; (ii) interdigitated (zipper-like) duplex structures form instead when sheared G·A base pairs are separated by one or two pairs of purine·purine mismatches; (iii) stacking is not restricted to base, deoxyribose also exhibits the potential to do so; (iv) canonical G·C or A·T base pairs are flexible enough to exhibit considerable changes from the regular H-bonded conformation. The paired bases become stacked when bracketed by sheared G·A base pairs, or become extruded out and perpendicular to their neighboring bases in the presence of interacting drugs; (v) the purine-rich and pyrimidine-rich loop structures are notably different in nature. The purine-rich loops form compact triloop structures closed by a sheared G·A, A·A, A·C or sheared-like Ganti·Csyn base pair that is stacked by a single residue. On the other hand, the pyrimidine-rich loops with a thymidine in the first position exhibit no base pairing but are characterized by the folding of the thymidine residue into the minor groove to form a compact loop structure. Identification of such diverse duplex or hairpin motifs greatly enlarges the repertoire for unusual DNA structural formation.  相似文献   

15.
Background and Aims The C4 perennial grass miscanthus has been found to be less sensitive to cold than most other C4 species, but still emerges later in spring than C3 species. Genotypic differences in miscanthus were investigated to identify genotypes with a high cold tolerance at low temperatures and quick recovery upon rising temperatures to enable them to exploit the early growing season in maritime cold climates. Suitable methods for field screening of cold tolerance in miscanthus were also identified.Methods Fourteen genotypes of M. sacchariflorus, M. sinensis, M. tinctorius and M. × giganteus were selected and grown under warm (24 °C) and cold (14 °C) conditions in a controlled environment. Dark-adapted chlorophyll fluorescence, specific leaf area (SLA) and net photosynthetic rate at a photosynthetically active radiation (PAR) of 1000 μmol m–2 s–1 (A1000) were measured. Photosynthetic light and CO2 response curves were obtained from 11 of the genotypes, and shoot growth rate was measured under field conditions.Key Results A positive linear relationship was found between SLA and light-saturated photosynthesis (Asat) across genotypes, and also between shoot growth rate under cool field conditions and A1000 at 14 °C in a climate chamber. When lowering the temperature from 24 to 14 °C, one M. sacchariflorus exhibited significantly higher Asat and maximum photosynthetic rate in the CO2 response curve (Vmax) than other genotypes at 14 °C, except M. × giganteus ‘Hornum’. Several genotypes returned to their pre-chilling A1000 values when the temperature was increased to 24 °C after 24 d growth at 14 °C.Conclusions One M. sacchariflorus genotype had similar or higher photosynthetic capacity than M. × giganteus, and may be used for cultivation together with M. × giganteus or for breeding new interspecies hybrids with improved traits for temperate climates. Two easily measured variables, SLA and shoot growth rate, may be useful for genotype screening of productivity and cold tolerance.  相似文献   

16.
Type 2 diabetes mellitus (T2DM) is associated with increased plasma triglyceride (TG) concentrations, but African Americans (AA) have lower plasma TG than Caucasians (CC). We evaluated the hypothesis that obese AA women have lower plasma TG than obese CC women do because of differences in lipid kinetics. Eleven AA and 11 CC obese women with T2DM, matched on body mass index (BMI) (AA = 37 ± 1, CC = 37 ± 1 kg/m2), age, duration of diabetes, percentage body fat, and insulin sensitivity (SI, determined by an intravenous glucose tolerance test), were studied. Plasma TG concentration (AA = 1.14 ± 0.11, CC = 1.88 ± 0.18 mmol/l), FFA rate of appearance (Ra) into plasma (AA = 419 ± 27, CC = 503 ± 31 µmol·min−1), and total VLDL-TG secretion rate (AA = 18 ± 2, CC = 29 ± 4 µmol·min−1) were lower in AA than CC women (all P < 0.05). In contrast, plasma total apolipoprotein (apo)B-100 concentration (AA = 1,542 ± 179, CC = 1,620 ± 118 nmol/l) and VLDL-apoB-100 secretion rate (AA = 1.3 ± 0.1, CC = 1.3 ± 0.1 nmol·min−1) were similar in both groups, so the molar ratio of VLDL-TG secretion rate to VLDL-apoB-100 secretion rate was lower in AA women than in CC women. VLDL-TG concentration was lower in AA women due to lower total VLDL-TG secretion rate. However, the VLDL-apoB-100 secretion rate was the same in both groups, demonstrating that AA women secrete smaller VLDL particles containing less TG than do CC women.  相似文献   

17.
Five strains (1126-1H-08T, 51B-09, 986-08, 1084B-08 and 424-08) were isolated from diseased rainbow trout. Cells were Gram-negative rods, 0.7 µm wide and 3 µm long, non-endospore-forming, catalase and oxidase positive. Colonies were circular, yellow-pigmented, smooth and entire on TGE agar after 72 hours incubation at 25°C. They grew in a temperature range between 15°C to 30°C, but they did not grow at 37°Cor 42°C. Based on 16S rRNA gene sequence analysis, the isolates belonged to the genus Flavobacterium. Strain 1126-1H-08T exhibited the highest levels of similarity with Flavobacterium oncorhynchi CECT 7678T and Flavobacterium pectinovorum DSM 6368T (98.5% and 97.9% sequence similarity, respectively). DNA–DNA hybridization values were 87 to 99% among the five isolates and ranged from 21 to 48% between strain 1126-1H-08T, selected as a representative isolate, and the type strains of Flavobacterium oncorhynchi CECT 7678T and other phylogenetic related Flavobacterium species. The DNA G+C content of strain 1126-1H-08T was 33.2 mol%. The predominant respiratory quinone was MK-6 and the major fatty acids were iso-C15∶0 and C15∶0. These data were similar to those reported for Flavobacterium species. Several physiological and biochemical tests differentiated the novel bacterial strains from related Flavobacterium species. Phylogenetic, genetic and phenotypic data indicate that these strains represent a new species of the genus Flavobacterium, for which the name Flavobacterium plurextorum sp. nov. was proposed. The type strain is 1126-1H-08T ( = CECT 7844T = CCUG 60112T).  相似文献   

18.
8-Oxoguanine (8-oxoG), induced by reactive oxygen species and arguably one of the most important mutagenic DNA lesions, is prone to further oxidation. Its one-electron oxidation products include potentially mutagenic guanidinohydantoin (Gh) and spiroiminodihydantoin (Sp) because of their mispairing with A or G. All three oxidized base-specific DNA glycosylases of Escherichia coli, namely endonuclease III (Nth), 8-oxoG-DNA glycosylase (MutM) and endonuclease VIII (Nei), excise Gh and Sp, when paired with C or G in DNA, although Nth is less active than the other two. MutM prefers Sp and Gh paired with C (kcat/Km of 0.24–0.26 min–1 nM–1), while Nei prefers G over C as the complementary base (kcat/Km 0.15–0.17 min–1 nM–1). However, only Nei efficiently excises these paired with A. MutY, a 8-oxoG·A(G)-specific A(G)-DNA glycosylase, is inactive with Gh(Sp)·A/G-containing duplex oligonucleotide, in spite of specific affinity. It inhibits excision of lesions by MutM from the Gh·G or Sp·G pair, but not from Gh·C and Sp·C pairs. In contrast, MutY does not significantly inhibit Nei for any Gh(Sp) base pair. These results suggest a protective function for MutY in preventing mutation as a result of A (G) incorporation opposite Gh(Sp) during DNA replication.  相似文献   

19.
We have cloned the M and S genes of the restriction-modification (R-M) system AhdI and have purified the resulting methyltransferase to homogeneity. M.AhdI is found to form a 170 kDa tetrameric enzyme having a subunit stoichiometry M2S2 (where the M and S subunits are responsible for methylation and DNA sequence specificity, respectively). Sedimentation equilibrium experiments show that the tetrameric enzyme dissociates to form a heterodimer at low concentration, with Kd ≈ 2 µM. The intact (tetrameric) enzyme binds specifically to a 30 bp DNA duplex containing the AhdI recognition sequence GACN5GTC with high affinity (Kd ≈ 50 nM), but at low enzyme concentration the DNA binding activity is governed by the dissociation of the tetramer into dimers, leading to a sigmoidal DNA binding curve. In contrast, only non-specific binding is observed if the duplex lacks the recognition sequence. Methylation activity of the purified enzyme was assessed by its ability to prevent restriction by the cognate endonuclease. The subunit structure of the M.AhdI methyltransferase resembles that of type I MTases, in contrast to the R.AhdI endonuclease which is typical of type II systems. AhdI appears to be a novel R-M system with properties intermediate between simple type II systems and more complex type I systems, and may represent an intermediate in the evolution of R-M systems.  相似文献   

20.
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   

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-23

京公网安备 11010802026262号