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.
O6-Methyl[8-3H]deoxyguanosine in a synthetic DNA polymer, poly(dC, dG, m6dG), is demethylated by cell-free extracts of EscherichiacoliBr adapted by exposure to N-methyl-N′-nitro-N-nitrosoguanidine, as shown by the appearance of 3H-labeled deoxyguanosine in hydrolysates of the recovered DNA. The demethylating activity could not be detected in extracts of nonadapted E. coli. These results provide direct evidence that a previously described inducible repair activity in E. coli acts by demethylating O6-methylguanine at the DNA level.  相似文献   

2.
By the use of space-filling models, a novel compound, 6-carbamylmethyl-8-methyl-7H-cyclopenta[f]isoquinolin-3(2H)-one (1) was devised which would be expected to hydrogen bond specifically to GC pairs in the major groove of the double helix such that (i) the amino group of the cytosine molecule donates a hydrogen bond to the C-3 carbonyl of the isoquinoline moiety and (ii) the amide proton of the side chain donates a hydrogen bond to the N-7 of guanine. From difference spectra studies it was found that 1 binds to native calf thymus DNA better than to denatured DNA; 1 inhibited RNA synthesis by a DNA-dependent RNA polymerase; and equilibrium dialysis experiments revealed that 1 binds to poly(dG).poly(dC), whereas no such binding to poly(dA).poly(dT) was observed.  相似文献   

3.
A single peak of DNA polymerase activity from extracts of T.brucei, obtained by DEAE-cellulose and phosphocellulose ion-exchange chromatography, was resolved into two peaks differing in KCl concentration necessary to elute them from a DNA-agarose column. Peak I (eluting at 0.2 M KCl) and Peak II (eluting at 0.4 M KCl), differed in response to increasing KCl concentrations, although both functioned optimally with Mg2+ as divalent cation when DNA synthesis was directed either by activated DNA or poly (dC)·(dG)12–18. Due to the potential significance of polyamines in the metabolism of T.brucei, the effect of exogenous polyamine on rates of DNA synthesis by the peak I and II enzymes was compared with that of murine DNA polymerase alpha. Only the peak I enzyme was significantly stimulated (up to 4-fold) by the biologically active polyamines spermine and spermidine at physiological concentrations. The response of the peak I enzyme resembled that of the alpha polymerase. This result suggests a possible functional difference between peak I and II enzymes, as well as a potential target site for trypanocidal drug development.  相似文献   

4.
The interactions between Escherichia coli RNA polymerase holoenzyme and a 3800 base-pair restriction fragment of bacteriophage T7 DNA (Mbo-IC) have been examined by electron microscopy. In addition to exhibiting weak, non-specific interactions (Ka ~ 104 M?1), RNA polymerase is able to form up to 15 to 20 relatively stable complexes with this template (Ka est 109 M?1). Only one of these complexes is formed at the T7 promoter E, that maps at 92.2 ± 1% on the conventionalgenome. The remaining complexes seem to be situated non-randomly on this fragment and possibly involve interactions with specific DNA sequences. The association kinetics of formation have been examined and give rise to a second-order rate constant of ~ 105 M?1s?1. Formation of these complexes is markedly reduced at low temperatures. Under standard binding conditions (50 mM-NaCl) the dissociation rate of these complexes is slow (t12 ~ 30 min), but increases rapidly with increasing salt concentration and at reduced temperatures. It is unaffected by the presence of heparin up to 5 μg/ml. Thus it appears that E. coli RNA polymerase can form complexes with promoter-like properties at many different sites on T7 DNA.  相似文献   

5.
Messenger RNA was isolated from rat preputial glands by guanidine HCl extraction, ethanol and salt precipitation, followed by chromatography on oligo(dT) cellulose. Double-stranded cDNA was synthesized from the mRNA and inserted into the Pst 1 site of the plasmid pBR322 by the poly(dG)·poly(dC) tailing and annealing procedure. The hybrid plasmids were used to transform E. coli HB101. Recombinant clones were screened for those containing cDNA inserts complementary to β-glucuronidase mRNA by a hybridization-selection procedure. One clone, containing an insert of about 1.2 kilobases, hybridized to preputial gland mRNA which, when translated in vitro, gave a product that migrated with the β-glucuronidase subunit on polyacrylamide gels.  相似文献   

6.
In a previous study, various intermediates in λ DNA packaging were visualized after lysis of λ-infected cells with osmotic shock and sedimentation through a sucrose formalin cushion onto electron microscope grids. Along this line, a systematic screening for intermediates accumulated in all head mutants available was performed. λA?-infected cells accumulate only empty spherical protein shells (petit λ) bound at an intermediate point along the DNA thread. In situ digestion experiments with restriction endonuclease EcoRI show that the petit λ-DNA complexes are formed at a fixed point on the DNA concatemer. In λNu1?-infected cells, however, most petit λ was not bound to DNA. In Fec? cells, which are defective in formation of concatemers but normal in head protein synthesis, most petit λ did not sediment onto the carbon film of the grid. In D? mutant, petit λ, partially full heads and empty heads with released DNA were observed. λFI?-infected cells also accumulate petit λ and partially full heads. The present studies suggest that protein pNu1 is required for complex formation between head precursors and DNA concatemers, pA for the initiation of DNA packaging, pD and pFI for the promotion of DNA packaging, and pD for stabilization of head structures. The results obtained with other head mutants involved in formation of mature proheads and head completion confirm earlier results obtained by different techniques.  相似文献   

7.
The number of silk fibroin genes per genome in the silkworm Bombyx mori has been determined by hybridization using fibroin [125I]mRNA. The purified [125I]mRNA had an oligonucleotide pattern after RNAase T1 digestion which was characteristic of fibroin mRNA (Suzuki &; Brown, 1972) and it hybridized specifically to DNA with a G + C content expected for a fibroin gene. Thermal denaturations indicated that these hybrids were mismatched by about 3%, which probably indicates some variation among the sequences encoding the internal repetitions of the fibroin protein.The concentration of fibroin gene sequences in B. mori DNA was measured by saturation hybridization of [125I]mRNA to filter bound DNA. The same saturation level of 1.8 × 10?5 μg mRNA per μg DNA was calculated from data obtained with unfractionated DNA and with fibroin gene sequences which had been separated from bulk B. mori DNA by actinomycin DCsCl centrifugation. Scatchard plots of the subsaturation data extrapolated to an identical saturation value. Internal reiteration of the fibroin mRNA molecule was apparent from the high association constant of hybridization. An exhaustive hybridization experiment showed that such repetitions comprise at least 90% of each mRNA molecule. The saturation value, in conjunction with the genome DNA content and the mRNA size, indicated the presence of only one fibroin gene per haploid B. mori genome.Hybridization of actinomycin DCsCl fractionated DNA indicated that fibroin mRNA can form hybrids with DNA that bands with bulk B. mori DNA. These hybrids appear to involve DNA which is related to, but distinguishable from, true fibroin gene sequences. The fibroin gene-related sequences form mismatched hybrids with the mRNA, are much shorter than the fibroin gene and are dispersed in B. mori DNA of much lower G + C content, and there are many copies of these sequences per B. mori genome.  相似文献   

8.
Neomycin inhibits in vitro DNA dependent DNA and RNA synthesis catalyzed by DNA polymerase I and RNA polymerase from E. coli. The effect of the antibiotic is more pronounced towards DNA synthesis. The inhibition of DNA synthesis is competitive with template DNA, does not reverse with excess deoxynucleoside triphosphate, Mg2+ or enzyme E. coli DNA polymerase I. Neomycin does not reduce the number of potential 3′ -OH end or primer. It seems to shorten the size of the newly formed polynucleotide.  相似文献   

9.
10.
Deamino-6-carba-oxytoxin (dC6O), a potent oxytocin analog considered to be resistant to some of the physiologically significant enzymic systems, and N-α-acetyl-[2-O-methyltyrosine]oxytocin (AMTO), an analog acting as a competitive inhibitor of oxytocin on the rat uterus, were studied in rats trained in a passive avoidance task.Subcutanaeous administartion of dC6O (5–50 gmg·kg?1) during different phases of the passive avoidance learning paradigm attenuated avoidance latencies; the results indicated that the drug induced state-dependent learning.AMTO (5–20 gmg·kg?1) enhanced avoidance latencies when administered subcutaneously before training trials and/or before retention test trials. This effect occured in both males and females. The analogs did not influence exploratory behavior in open field.The results suggest that oxytoxin, in contrast to vasopressin, may impair memory processes. However, both analogs failed to influence the passive avoidance response when administered after training. This finding indicates that dC6O and AMTO did not influence the mechanism of memory consolidation whereas vasopressin and oxytoxin had a marked effect.  相似文献   

11.
Treatment of the eukaryotic organism Tetrahymena pyriformis with low concentrations of Ethidium Bromide causes accumulation of a protein-nucleic acid complex consisting of a DNA polymerase, a RNA polymerase, a deoxyribo-nuclease and a RNA linked DNA fragment. The length of the RNA is about 30 nucleotides, while the DNA part is around 200 nucleotides long. Degradations with ribonucleases and deoxyribonucleases strongly indicate that the RNA exists in a non-hybrid structure with a homogenous base composition and that the DNA is single-stranded. The complex is purified 1100 fold from whole cells and sodium dodecyl sulphate acrylamide gel electrophoresis gives 9 defined bands. The polynucleotide in the isolated complex accounts for only 10?4 of the total cellular DNA.As the complex contains some of the enzymes essential for discontinous DNA replication, in addition to a RNA linked Okazaki fragment, it is concluded that a highly purified replication complex has been isolated.  相似文献   

12.
The magnesium ion-dependent equilibrium of vacant ribosome couples with their subunits
70 S?k?1k150 S+30S
has been studied quantitatively with a novel equilibrium displacement labeling method which is more sensitive and precise than light-scattering. At a concentration of 10?7m, tight couples (ribosomes most active in protein synthesis) dissociate between 1 and 3 mm-Mg2+ at 37 °C with a 50% point at 1.9 mm. The corresponding association constants Ka′ are 5.1 × 105m?1 (1 mm-Mg2+), 3.5 × 107m?1 (2 mm), and 1.2 × 109m?1 (3 mm), about five orders of magnitude higher than the Ka′ value of loose couples studied by Spirin et al. (1971) and Zitomer & Flaks (1972).In this range of Mg2+ concentrations (37 °C, 50 mm-NH4+) the rate constants depend exponentially and in opposite ways on the Mg2+ concentration: k1 = 2.2 × 10?3s?1, k?1 = 7.7 × 104m?1s?1 (2mm-Mg2+); k1 = 1.5 × 10?4s?1, k?1 = 1.7 × 107m?1s?1 (5 mm-Mg2+). Under physiological conditions (Mg2+ ~- 4 mm, ribosome concn ~- 10?7m), the equilibrium strongly favors association and the rate of exchange is slow (t12 ~- 10 min). In the range of dissociation (2 mm-Mg2+), association of subunits proceeds without measurable entropy change and hence ΔGO = ΔHO. The negative enthalpy change of ΔHO = ? 10 kcal suggests that association of subunits involves a shape change.Below a critical Mg2+ concentration (~- 2 mm), the 50 S subunits are converted irreversibly into the b-form responsible for the transition to loose couples. The results are compatible with two classes of binding sites, one class binding Mg2+ non-co-operatively and contributing to the free energy of association by reduction of electrostatic repulsion, and another class probably consisting of hydrogen bonds between components at opposite interfaces whose critical spatial alignment rapidly denatures in the absence of stabilizing magnesium ions.  相似文献   

13.
5-Fluoroorotic acid treatment lowered the (Guanine + Cytosine)/(Adenine + Uracil) base ratio of 32P-labeled microsomal RNA from a control value of 1.36 to 1.00. Low doses of actinomycin D, which are effective in inhibiting ribosomal RNA synthesis without significantly affecting messenger RNA synthesis, caused a similar decrease in the base ratio. Microsomal RNA labeled by [3H]orotate in the presence of 5-fluoroorotic acid had approximately 12 the specific radioactivity but twice the hybridization efficiency of RNA labeled in its absence. Evidence is presented that this RNA (1) has a different structure from that of ribosomal RNA, (2) hybridizes to DNA with an efficiency consistent with that of other published studies of polysome-associated messenger RNA, and (3) possesses sequences which are present in other samples of liver microsomal RNA but not in kidney microsomal RNA. These properties differ from those known to be exhibited by 18 S and 28 S ribosomal RNA. Electrophoretic analysis of this [3H]orotate-labeled microsomal RNA indicated that the analogue greatly inhibited precursor incorporation into ribosomal RNA but had little or no effect on incorporation into messenger RNA. Ribosomal RNA and polyadenylate-rich nonribosomal RNA were prepared from total polyribosomes by phenol extraction at pH 7.6 and pH 9.0, respectively. 5-Fluoroorotic acid inhibited [3H]orotate or 32Pi incorporation into the pH 7.6 fraction much more effectively than incorporation into the pH 9.0 fraction. A subfraction of the pH 9.0 RNA which was retained by a polythymidylate-cellulose column had a greatly increased adenylate content.  相似文献   

14.
This report presents an analysis of histone gene expression in the cleaving embryo of the sea urchin, Strongylocentrotus purpuratus, with emphasis on whether the regulatory site(s) in the pathway of gene expression change as development proceeds. The analysis focuses on the equation, dP1dt = M·f·n·At, where dP1dt = the absolute rate of histone synthesis; M = the mole quantity of histone messenger RNA; f = the fraction of histone mRNA in polysomes; n = the polysome size; and At = the rate of elongation of nascent histone polypeptide chains. The embryo solves this rate equation differently at different times. Measurements were made (at 15°C) of absolute rates of histone synthesis (dP1dt). The rate of histone synthesis increases at least 48-fold during the first 6 hr after fertilization from less than 0.5 to 24 pg embryo?1 hr?1; in the period from 6 to 12 hr, this rate rises to 182 pg embryo?1 hr?1, an additional 7.7-fold rise, resulting in an overall increase of 370-fold between the 1-cell and 200-cell stage. The fraction of newly synthesized (zygotic) histone messenger RNA that partitions into polysomes (fzygotic) has also been measured during the first 12 hr of development. This fraction increases from 0.2 in the 2-hr embryo to 0.8 in the 6-hr embryo (16-cell stage), increasing slowly thereafter to near unity by 12 hr. The size of histone-synthesizing polysomes (n) does not change substantially over the 12-hr interval, remaining constant at a weighted mean of 5 ribosomes per polysome (range 3 to 7). Utilizing the data on fzygotic and dP1dt, the rate of elongation of nascent histone polypeptide chains (At) during the first 6 hr of development was estimated; At remains constant at 1.11 codons per second. This calculated value is in fair agreement with a direct measurement of histone peptide elongation rate in the 12-hr embryo. It is proposed that histone gene expression in cleaving sea urchin embryos be divided into two phases, distinguished on the basis of their pivotal translational parameters: Phase I (0–6 hr), during which f is rate determining, and Phase II (6 hr on), during which M is the rate-determining parameter.  相似文献   

15.
The binding site and the geometry of Co(III)meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (CoTMPyP) complexed with double helical poly(dA)·poly(dT) and poly(dG)·poly(dC), and with triple helical poly(dA)·[poly(dT)]2 and poly(dC)·poly(dG)·poly(dC)+ were investigated by circular and linear dichroism (CD and LD). The appearance of monomeric positive CD at a low [porphyrin]/[DNA] ratio and bisignate CD at a high ratio of the CoTMPyP-poly(dA)·poly(dT) complex is almost identical with its triplex counterpart. Similarity in the CD spectra was also observed for the CoTMPyP-poly(dG)·poly(dC) and -poly(dC)·poly(dG)·poly(dC)+ complex. This observation indicates that both monomeric binding and stacking of CoTMPyP to these polynucleotides occur at the minor groove. However, different binding geometry of CoTMPyP, when bind to AT- and GC-rich polynucleotide, was observed by LD spectrum. The difference in the binding geometry may be attributed to the difference in the interaction between polynucleotides and CoTMPyP: in the GC polynucleotide case, amine group protrude into the minor groove while it is not present in the AT polynucleotide.  相似文献   

16.
tRNACUGLeu from E. Coli was purified by column chromatography on benzoylated DEAE-cellulose, followed by hydroxyapatite prepared by an improved method. Crystals obtained by vapour diffusion gave X-ray diffraction out to 7 Å in the hk0 projection and 10 Å in h0?. The space group was P42212 with a = b = 133 A?, c = 66 A? and 8 molecules in the unit cell. Birefringence showed preferred orientation of RNA helical regions in the ab plane.  相似文献   

17.
Membrane vesicles from pigeon erythrocytes show a rapid, ATP-dependent accumulation of 45Ca2+. Ca2+ accumulation ratios greater than or approximately equal to 104 are readily attained. For ATP-dependent Ca2+ uptake, V is 1.5 mmol · 1?1 · min?1 at 27°C (approx. 0.9 nmol · mg?1 protein · min?1), [Ca2+]12 is 0.18 μM, [ATP]12 is 30–60 μM, the Ca2+ uptake rate depends on [Ca2+]2 and the dependence of uptake rate on ATP concentration implies strong ATP-ATP cooperativity. The Arrhenius activation energy is 19.1 ± 1.4 kcal/mol and the pH optimum is approx. 6.9.  相似文献   

18.
Ryegrass, harvested at the pre-ear emergence stage of growth, was ensiled in laboratory silos, either fresh (175 g dry matter kg?1) or wilted to five DM levels ranging from 216–432 g DM kg?1, with and without additive treatment. The additives used were “Sylade” containing sulphuric acid (15%) and formaldehyde (23%) applied at 4.6 l t?1 and an “ADD-F” (85% formic acid)formalin mixture (7:3 by volume) applied at a similar rate (4.8 l t?1). An additional treatment included application of the mixture at a constant rate related to the DM content of the ensiled crop (25 l t?1 DM).In the untreated silages, the water-soluble carbohydrates (WSC) varied, respectively (over the DM range 175–432), from 0–32 g kg?1 DM compared with 197-6 g kg?1 DM for the “Sylade” treated silages and 256-50 g kg?1 DM for the formic acid/formalin silages treated at an additive rate of 4.8 l t?1. Corresponding ranges of protein N for the control and treatments (expressed as g kg?1 total N) were 302–447, 624-502 and 620-505, respectively. When the formic acid/formalin additive was applied at a constant level related to the DM content of the crop, although the WSC content decreased with increasing DM (247-158 g kg?1 DM), the protein N content (612 g kg?1 total N) remained constant.Grass from the same field was ensiled fresh, treated with “ADD-F” at the rate of 3.4 l t?1 fresh grass, ADD-Fformalin at the rate of 4.8 l t?1 fresh grass and “Sylade” at the rate of 4.6 l t?1 fresh grass. The silages were given to Suffolk-cross wether lambs in digestibility and intake trials. Digestibility coefficients of DM and energy of the silage treated with “Sylade” were significantly lower (P < 0.05) than those of the other three silages. The DM intakes of all the silages were high, ranging from 27.7 g kg?1 live weight for the “Sylade” silage to 30.7 g kg?1 live weight for the silage treated with ADD-Fformalin. Live weight gains ranged from 200 g/day for the control silage to 267 g/day for the ADD-Fformalin silage.  相似文献   

19.
Studies were carried out to determine the Hill coefficients for the inhibition by F? of the erythrocyte membrane-bound Mg2+-ATPase, (Na+ + K+)-ATPase and acetylcholinesterase from rats fed with seven different diets. Five groups were fed with different natural fats or oil supplements, one with a hydrogenated fat supplement and the other with fat-free diet. The responses of the red cell fatty acids to dietary fats were recorded. The value of n for the inhibition by F? of the three enzymes revealed a particular and different behaviour in each group. Correlations between the fatty acid compositions of erythrocyte membranes and cooperativity of each enzyme were calculated. The results indicate that neither the essential fatty acid family nor the non-essential ones are particularly involved in the allosteric phenomena. The increase of the double bond index/saturation ratio of fatty acids, which is taken as indicative of membrane fluidity, was accompanied in an inverse manner by changes in allosteric transitions of the (Na+ + K+)-ATPase and acetylcholinesterase, whereas the Mg2+-ATPase was not dependent on this ratio. Diminution of membrane fluidity, carried out by in vitro increase of its cholesterol content, yields confirmatory results of this regulatory mechanism since the value of n for acetylcholinesterase shifted as predicted.These facts indicate that the membrane fluidity is a physiological regulator for the allosteric behaviour of the membrane-bound enzymes and that each enzyme exhibits a particular behaviour in this phenomenon.  相似文献   

20.
Four marine dinoflagellates, Amphidinium carterae Hulburt, Ceratium tripos (O.F. Müll.) Nitzsch, Prorocentrum minimum (Pav.) J. Schiller, and Scrippsiella trochoidea (Stein) Loeblich III were grown as dilution cultures at 18°C, S = 29%. and 30 μE·m?2·s?1 at L:D = 14:10 h. In nutrient-saturated cultures, the growth rates (doubl·day?1) ranged from 0.38 for Scrippsiella to 0.80 for Prorocentrum, and carbon content (pg·cell?1) from 83 for Amphidinium to 6900 for Ceratium. The atomic NC ratio was 0.13–0.15, but for Ceratium it was 0.088, because of its thick, cellulose theca. The atomic NP ratio ranged from 12–13 for Ceratium and Scrippsiella to 15–17 for Prorocentrum and Amphidinium. Under P-deficient conditions (growth rate 39–70% of the maximum), cellular P decreased considerably, but so did N, so that the NP ratio was only slightly affected. There was a concomitant increase in carbon content per cell of 1.2- to 1.7-fold. Alkaline phosphatase activity was virtually nil in nutrient-saturated cells, but was readily demonstrable in all species when P-deficient.  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