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.
2.
The complexity of Rous sarcoma virus RNA has been determined using molecular hybridization. Relative to poliovirus RNA, the complexity of Rous sarcoma virus is 9·3 × 106 daltons, a value close to its physically-determined molecular weight of about 107. Our interpretation is that the 35 S RNA subunits of the 70 S virus genome are non-repetitive, that is, each possesses a unique nucleotide sequence, although a limited amount of redundancy cannot be excluded.  相似文献   

3.
The mitochondrial genome of wild-type yeast cells. IV. Genes and spacers   总被引:12,自引:0,他引:12  
The organization of the mitochondrial genome of wild-type Saccharomyces cerevisiae cells has been investigated further, by degrading mitochondrial DNA with micrococcal nuclease. Under the conditions used, this enzyme very strongly degrades the A + T-rich stretches (spacers) whereas it only inflicts a limited number of breaks into the G + C-rich stretches (genes). The macromolecular fragments derived from the “genes” have been separated from the oligonucleotides originating from the “spacers” by gel filtration, and both sorts of products have been investigated. It has been shown (a) that the spacers are very homogeneous in base composition and have a G + C content lower than 5% (mitochondrial DNA has a G + C content of 18%); (b) that the genes are very heterogeneous in base composition, the G + C content ranging from about 25% to 50%, when the average size of the fragments is 1·2 × 105; smaller fragments, molecular weight 4 × 104, having a G + C level as high as 65%, have been isolated in a yield of 10%; the average G + C content of genes is about 32%; (c) that genes and spacers are present in about equal amounts in the mitochondrial genome and that they have comparable average sizes.  相似文献   

4.
In the rat phrenic nerve-diaphragm muscle preparation, X-537A at 6×10?6 to 3×10?5 M (1) depolarized muscle fibre membranes, (2) caused an occasional transient increase in and ultimate block of spontaneous transmitter release, (3) did not increase the amplitude of the end-plate potential (epp) but abruptly blocked stimulus-evoked transmitter release, and (4) produced an increase in the occurrence of “giant” miniature epp's (mepp's). The possibility is discussed that the sporadically raised mepp frequency was due to an ionophore-induced depolarization of nerve terminals. The increased occurrence of “giant” mepp's apparently reflected a X-537A-induced spontaneous multiquantal release of acetylcholine. This was not dependent on extracellular calcium but appeared to be of presynaptic origin.  相似文献   

5.
The effect of molecular “aging” of fibrinogen stimulated by preincubation in solution on the fibrin three-dimensional architecture, its ability to crosslink fibrin-stabilizing factor, and the sensitivity of fibringel to plasmin hydrolysis have been studied. The method of elastic light scattering was used to demonstrate that fibrin generated from “defective” fibrinogen had a coarser structure with a higher mean mass-length ratio of polymeric fibers compared to native fibrinogen (2.24 × 109 and 1.46 × 109 g/(mol cm), respectively). Crosslinking had no effect on the architecture of both control and experimental fibrin samples. Spectrophotometric and electrophoretic analysis has shown a higher sensitivity of coarse fibrin gels to plasmin. A close correlation between spontaneous local conformational reconstructions in fibrinogen molecule and its functional activity is concluded.  相似文献   

6.
Analysis of the binding of phenylalanine to phenylalanyl-tRNA synthetase   总被引:1,自引:0,他引:1  
Using the complete rate equation for the PPi-ATP exchange reaction at equilibrium, the dissociation constants of phenylalanine (10?5m), phenylalanine butyl ester (8 × 10?5m), benzyl alcohol (6 × 10?4m), phenylalaninol (2 × 10?4m), hydrocinnamic acid (3 × 10?3m) and glycine (>1 m) with the phenylalanyl-tRNA synthetase (Escherichia coli K12) were determined. Taking the model of Koshland (1962) for the estimation of the configurational free energy change due to proximity and orientation, and decomposing the process of binding into several thermodynamic steps, the contribution to binding of the benzyl group, glycine unit, protonated amino group, carboxylate group and joint interactions were estimated. The results are: (1) the standard free energy contributions for binding phenylalanine are benzyl group (?8.2 kcal/mol), glycine unit (?2.5 kcal/mol), protonated amino group (?0.8 kcal/mol) and carboxylate group (1 kcal/mol). (2) The standard free energy change due to the change in the interaction between the protonated amino group and carboxylate group when they are transferred from the aqueous environment to the enzyme environment is ?2.7 kcal/mol. (3) A dissociation constant for glycine of 7.5 m is calculated without the hypothesis that a conformational change occurs in the enzyme when the benzyl unit of phenylalanine binds, permitting an interaction of the enzyme with the protonated amino and/or carboxylate groups.The detection of E·AA2 and E·ATP shows that a sequential addition of substrates is not necessary for binding. A comparison of the dissociation constants of E·AA (10?5m), E·ATP (1.5 × 10?3m), E·PP (5.5 × 10?4m), E·I (8 × 10?5m) and the mixed complexes E·I·ATP (6 × 10?8m2), E·I·PP (5 × 10?8m2) and E·AA·PP (7 × 10?9m2), with phenylalanine butyl ester as the inhibitor, indicates no strong interaction between the binding of ATP or PPi with the binding of phenylalanine.  相似文献   

7.
The unusual nucleotide guanosine tetraphosphate, ppGpp, which appears following amino acid starvation in “stringent” strains of bacteria binds to the elongation factor EFTu with a dissociation constant of about 8 × 10?9m. ppGpp binds competitively with GDP and GTP, and EFTs catalyzes the exchange reaction of ppGpp with EFTu · GDP. ppGpp binds to EFTu about 50 times more tightly than does GTP, and, in the absence of elongation factor EFTs, it will effectively inhibit the formation of the ternary complex Phe-tRNA · EFTu · GTP. However, in the presence of EFTs there is rapid equilibration between EFTu · GTP and EFTu · ppGpp which allows EFTu to be rapidly and extensively incorporated into the stable ternary complex. A preliminary estimate of the constant for the dissociation of Phe-tRNA from the ternary complex is 10?810?9m. ppGpp inhibits the enzymatic binding of Phe-tRNA to ribosomes; however, EFTs reverses this inhibition. ppGpp moderately inhibits phenylalanine polymerization even in the presence of EFTs. This inhibition probably involves an interaction of ppGpp with elongation factor G, the translocation factor. It appears that in the intact cell ppGpp would not be an effective inhibitor of EFTu, and that little EFTu · ppGpp can exist in the cell.  相似文献   

8.
S R Weiss  H E Varmus  J M Bishop 《Cell》1977,12(4):983-992
The genome of avian sarcoma virus (ASV) contains four known genes: gag, encoding structural proteins of the viral core; pol, encoding the viral RNA-directed DNA polymerase; env, encoding the glycoprotein(s) of the viral envelope; and src, which is responsible for neoplastic transformation of the host cell. We have located these genes on virus-specific RNAs in cells productively infected with both nondefective and defective strains of ASV by using molecular hybridization with DNAs complementary to specific portions of the ASV genome.The cytoplasm of cells producing nondefective ASV contains three species of polyadenylated virus-specific RNA, each of which has chemical polarity identical to that of the viral genome. The largest species has a molecular weight of 3.3 × 106 daltons and a sedimentation coefficient of 38S, encodes all four viral genes, and is probably identical to the viral genome. A second species has a molecular weight of 1.8 × 106 daltons and a sedimentation coefficient of 28S, and encodes the 3′ half of the viral genome, including env, src and a genetically silent region known as “c.” The smallest species has a molecular weight of 1.2 × 106 daltons and a sedimentation coefficient of 21S, and encodes only src and “c.” All three species of virus-specific RNA contain nucleotide sequences at least partially homologous to a sequence of 101 nucleotides found at the extreme 5′ end of the ASV genome. This sequence may not be present in the portions of the ASV genome which encode the 28S and 21S virus-specific RNAs, and hence may be joined to these RNAs during their maturation from precursor molecules.The size and genetic composition of virus-specific RNAs in cells producing defective deletion mutants reflect the nature of the deletion. Deletions of either src or env eliminate the 28S virus-specific RNA, leaving a 21S RNA (which contains either env and “c” in the case of src deletions or src and “c” in the case of env deletions) and a 35S RNA which is probably identical to the viral genome.Based on these and related results, we propose a model for viral gene expression which conforms to previous suggestions that eucaryotic cells initiate translations only at the 5′ termini of messenger RNAs.  相似文献   

9.
Twenty-eight Bam H 1 restriction fragments were isolated from normal mitochondrial DNA of maize by recombinant DNA techniques to investigate the organization of the mitochondrial genome. Each cloned fragment was tested by molecular hybridization against a Bam digest of total mitochondrial DNA. Using Southern transfers, we identified the normal fragment of origin for d each clone. Twenty-three of the tested clones hybridized only to the fragment from which the clone was derived. In five cases, labeling of an additional band indicated some sequence repetition in the mitochondrial genome. Four clones from normal mitochondrial DNA were found which share sequences with the plasmid-like DNAs, S-1 and S-2, found in S male sterile cytoplasm. The total sequence complexity of the clones tested is 121×106 d (daltons), which approximates two thirds of the total mitochondrial genome (estimated at 183×106 d). Most fragments do not share homology with other fragments, and the total length of unique fragments exceeds that of the largest circular molecules observed. Therefore, the different size classes of circular molecules most likely represent genetically discrete chromosomes in a complex organelle genome. The variable abundance of different mitochondrial chromosomes is of special interest because it represents an unusual mechanism for the control of gene expression by regulation of gene copy number. This mechanism may play an important role in metabolism or biogenesis of mitochondria in the development of higher plants.  相似文献   

10.
The kinetics of the hydrogen-deuterium exchange reactions of double-helical poly (rI) · poly (rC), single-stranded poly(rC) and poly(rI), inosine, and cytosine- 5′-phosphoric acid have been examined, at various temperatures in the range 20 °C to 52 °C, by stopped-flow ultraviolet spectrophotometry, in the region 270 to 300 nm. For the solution of double-helical poly(rI) · poly(rC), two first-order deuteration reactions were found: a fast one and a slow one. At 25 °C and at pH 7.0, the rate constant was 12.3 s?1 for the fast reaction, and 0.13 s?1 for the slow reaction. The rate constant of the fast reaction is nearly equal to that of the single-stranded poly(rC) (12.6 s?1), and is assigned to the deuteration at the amino hydrogen (that is, free from the C · I hydrogen bond) of the cytosine residue. The slow reaction is attributable to the deuteration of the two hydrogens: the amino hydrogen of rC and imide hydrogen of rI, which are rapidly exchanging with each other within every rC · rI base-pair. From the observed temperature effect on this slow reaction rate, it has been concluded that there are two types of “opening process” that are relevant to the hydrogen exchange reaction; one of them is predominent in the range 47 °C to 52 °C and the other in the temperature region lower than 47 °C. The enthalpy (H) and entropy (S) differences of the “open” and “closed” forms in the former type process are ΔH = 167 kcal per mole and ΔS = 507 e.u., while in the latter ΔH = 8.1 kcal per mole and ΔS = 10 e.u..  相似文献   

11.
The distribution of living coccolithophores in the California Current system of southern California at 10 m water depth was investigated on two dates in March and June, 1982. Six closely spaced stations were sampled in March, of which three were resampled in June. Thirty-six euphotic species were identified of which four,Emiliania huxleyi, Umbilicosphaera sibogae, Gephyrocapsa oceanica, andRhabdosphaera longistylis, respectively, were the most abundant. Both the “cold” and “warm” morphotypes ofE. huxleyi were present, in varying proportions. Large ranges in community structure, diversity (0.35–2.64 natural bels), and standing crop (1.0 × 104–6.2 × 105cells/l) were recorded. This range of end-member values is approximately that found in the open ocean from 0° to about 65° latitude.The distributions of four coccolithophore assemblages recognized in March samples from the Borderland area appear to reflect the following distinct water masses: (1) California Current; (2) Southern California Counter Current; (3) Transitional Zone; (4) Near Shore. The coccolithophore assemblages from the June stations were more uniform, indicating that the Borderland was experiencing more stable conditions than in March.  相似文献   

12.
Human bone marrow colony growth in agar-gel   总被引:73,自引:0,他引:73  
A technique for growing human bone marrow cell colonies in agar-gel medium is described. “Feeder layers” containing 1 × 106 normal human peripheral white blood cells are used as the stimulus for colony growth. Human bone marrow aspirates are collected in heparinized syringes and plated as 2 × 105 cells on “feeder layers.” Normal human bone marrow yields 32–102 colonies per 2 × 105 cells plated. Colonies are almost exclusively granulocytic. Growth rate of colonies is slower than with mouse bone marrow but colonies reach a comparable size (500–1500 cells) at days 12–16.  相似文献   

13.
Sequence complexity of nuclear RNAs in adult rat tissues   总被引:26,自引:0,他引:26  
D M Chikaraishi  S S Deeb  N Sueoka 《Cell》1978,13(1):111-120
  相似文献   

14.
Here, I provide the first direct estimate of the spontaneous mutation rate in an Old World monkey, using a seven individual, three‐generation pedigree of African green monkeys. Eight de novo mutations were identified within ~1.5 Gbp of accessible genome, corresponding to an estimated point mutation rate of 0.94 × 10?8 per site per generation, suggesting an effective population size of ~12000 for the species. This estimation represents a significant improvement in our knowledge of the population genetics of the African green monkey, one of the most important nonhuman primate models in biomedical research. Furthermore, by comparing mutation rates in Old World monkeys with the only other direct estimates in primates to date–humans and chimpanzees–it is possible to uniquely address how mutation rates have evolved over longer time scales. While the estimated spontaneous mutation rate for African green monkeys is slightly lower than the rate of 1.2 × 10?8 per base pair per generation reported in chimpanzees, it is similar to the lower range of rates of 0.96 × 10?8–1.28 × 10?8 per base pair per generation recently estimated from whole genome pedigrees in humans. This result suggests a long‐term constraint on mutation rate that is quite different from similar evidence pertaining to recombination rate evolution in primates.  相似文献   

15.
NADPH-cytochrome c reductase has been isolated from a top-fermenting ale yeast, Saccharomyces cerevisiae (Narragansett strain), after ca. a 240-fold purification over the initial extract of an acetone powder, with a final specific activity (at pH 7.6, 30 °C) of ca. 150 μmol cytochrome c reduced min?1mg?1 protein. The preparation appears to be homogeneous by the criteria of: sedimentation velocity; electrophoresis on cellulose acetate in buffers above neutrality; and by polyacrylamide gel electrophoresis. Although the reductase appeared to partially separate into species “A” and “B” on DEAE-cellulose at pH 8.8, the two species have proven to be indistinguishable electrophoretically (above pH 8) and by sedimentation. By sedimentation equilibrium at 20 °C, a molecular weight of ca. 6.8 (± 0.4) × 104 was obtained with use of a V?20 ° = 0.741 calculated from its amino acid composition. After disruption in 4 m guanidinium chloride- 10 mm dithioerythritol- 1 mm EDTA, pH 6.4 at 20 °C, an M?r of 3.4 (± 0.1) × 104 resulted, which points to a subunit structure of two polypeptide chains per mole. Confirmatory evidence of the two-subunit structure with similar, if not identical, polypeptide chains was obtained by polyacrylamide gel electrophoresis in dodecyl-sulfate, after disruption in 4 m urea and 2% sodium dodecyl sulfate, and yielded a subunit molecular weight of ca. 4 × 104. Sulfhydryl group titration with 4,4′-dithiodipyridine under acidic conditions revealed one sulfhydryl group per monomer, which apparently is necessary for the catalytic reduction of cytochrome c. NADPH, as well as FAD, protects this-SH group from reaction with 5,5′-dithiobis (2-nitrobenzoate). The visible absorption spectrum of the oxidized enzyme (as prepared) has absorption maxima at 383 and 455 nm, typical of a flavoprotein. Flavin analysis (after dissociation by thermal denaturation of the “A” protein) conducted fluorometrically, revealed the presence of 2.0 mol of FAD per 70,000 g, in confirmation of the deduced subunit structure. The identity of the FAD dissociated from either “A” or “B” protein was confirmed by recombination with apo-d-amino acid oxidase and by thin-layer chromatography. A kinetic approach was used to estimate the dissociation constant for either FAD or FMN (which also yields a catalytically active enzyme) to the apoprotein reductase at 30 °C and pH 7.6 (0.05 m phosphate) and yielded values of 4.7 × 10?8m for FAD and 4.4 × 10?8m for FMN.  相似文献   

16.
Analysis by electrophoresis in polyacrylamide gels, followed by silver staining, of dsRNA extracted from many samples of raspberry leaves infected with raspberry leaf mottle virus (RLMV) and/or raspberry leaf spot virus (RLSV) failed to detect reliably any significant quantities of dsRNA species in excess of 1·0 × 106mol. wt. This contrasts with results reported from Canada where three dsRNA species of estimated mol. wt 2·6 × 1061·6 × 106and 1·1 × 106were consistently associated with infection with RLSV but none were associated with RLMV. However, in Scotland, four dsRNA species of estimated mol. wt 2·4 × 1061·6 × 1060·7 × 106and 0·3 × 106were detected in raspberry infected with apple mosaic ilarvirus. These results suggest that the dsRNA species reported from Canada are not those of RLSV but are probably those of a second virus, possibly an ilarvirus, which occurs together with RLSV and/or induces similar symptoms. A few samples from plants infected with RLMV and RLSV contained very small amounts of two dsRNA species of estimated mol. wt 4·7 × 106and 4·5 × 106. It is not known whether these species are those of RLMV and RLSV.  相似文献   

17.
N‐carbamoyl‐amino‐acid amidohydrolase (also known as N‐carbamoylase) is the stereospecific enzyme responsible for the chirality of the D ‐ or L ‐amino acid obtained in the “Hydantoinase Process.” This process is based on the dynamic kinetic resolution of D ,L ‐5‐monosubstituted hydantoins. In this work, we have demonstrated the capability of a recombinant L ‐N‐carbamoylase from the thermophilic bacterium Geobacillus stearothermophilus CECT43 (BsLcar) to hydrolyze N‐acetyl and N‐formyl‐L ‐amino acids as well as the known N‐carbamoyl‐L ‐amino acids, thus proving its substrate promiscuity. BsLcar showed faster hydrolysis for N‐formyl‐L ‐amino acids than for N‐carbamoyl and N‐acetyl‐L ‐derivatives, with a catalytic efficiency (kcat/Km) of 8.58 × 105, 1.83 × 104, and 1.78 × 103 (s?1 M?1), respectively, for the three precursors of L ‐methionine. Optimum reaction conditions for BsLcar, using the three N‐substituted‐L ‐methionine substrates, were 65°C and pH 7.5. In all three cases, the metal ions Co2+, Mn2+, and Ni2+ greatly enhanced BsLcar activity, whereas metal‐chelating agents inhibited it, showing that BsLcar is a metalloenzyme. The Co2+‐dependent activity profile of the enzyme showed no detectable inhibition at high metal ion concentrations. © 2010 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

18.
Different concentrations of in utero incubated rabbit sperm (1.5 × 104-120 × 104 /ml) were tested to determine whether there is a relationship between sperm concentration and level of fertilization achieved “in vitro” of rabbit ova. While low concentrations (1.5 × 104-4.5 × 104 /ml) resulted in relatively low fertilization (23–36%), those in the range of 13 × 104?120 × 104 /ml gave fertilization rates of 65–83%. Consistently high results were obtained with sperm counts above 40 × 104 /ml. This is in agreement with the concentration of spermatozoa found in vivo in the Fallopian tubes around the time of fertilization (50 × 104 /ml).  相似文献   

19.
Photosynthetic fructose-1,6-diphosphatase (FDPase) fractions I and II, earlier purified from spinach leaves, show a similar amino acid composition, with the exception of a higher glutamic acid content in the latter. In both fractions glutamic and aspartic acids are the main amino acids. pH activity profiles of fractions I and II are similar, with optima at 8·65–8·70, both showing a high specificity for fructose- 1,6-diphosphate. These two fractions are Mg2+-dependent for activity, with an Optimum Mg2+ concentration of 10 mM in standard conditions, which shifts to 5 mM when the MG2+/EDTA ratio is increased to 10; Mn2+ and Co2+ are slightly active. EDTA enhances FDPase activity slightly, with an optimum at 0·4–0·8 mM. Cysteine has no activating effect, and acts as an inhibitor above 10 mM. Both I and II have an optimum substrate concentration of 4 mM, and the substrate inhibits at concns above this value. Kinetic velocity curves are sigmoidal, with the concave zone located in the range of physiological substrate concns. (Hill coefficient 1·75 for both). This suggests a strong regulatory role of fructose-1,6-diphosphate. Km values are 1·4 × 10−3 M (fraction I) and 1·1 × 10−3 M (fraction II). The highest activity rate occurs at 60°, in accordance with the high thermostability of both fractions; the activation energies are 14·3 kcal/mol (fraction I) and 13·0 kcal/mol (fraction II).  相似文献   

20.
The transfecting efficiency of P22 DNA on “rough” strains of Salmonella typhimurium or non-restricting mutants of Escherichia coli K12 approaches 3 × 10?8 plaques/genome equivalent. It increases 20-fold upon complete erosion of the terminally redundant regions of the DNA molecule with either λ exonuclease or exonuclease III. Eroded DNA molecules form circles and linear oligomers upon annealing. The circular monomers display transfecting activity about ten times higher than that of eroded linear monomers or hydrogen-bonded oligomers. recB recC sbcB strains of E. coli K12 are transfected with P22 DNA with an efficiency of 1.5 × 10?6 plaques/genome equivalent. The activity of DNA molecules on these strains is not augmented by erosion. This suggests that the activation by erosion, seen in assays on rec+ genotypes, is due to the formation of hydrogen-bonded circular molecules, which more readily escape degradation by the recBC nuclease.  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