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.
The uptake of various nucleosides by S49 mouse T-lymphoma cells and that by a single-step nucleoside transport-defective mutant thereof (AE1) were compared. Residual nucleoside entry into AE1 cells occurred via two routes, nonmediated permeation and saturable, non-concentrative transport with broad substrate specificity and a Michaelis-Menten constant approximating that for thymidine transport in wild-type cells. However, in contrast to nucleoside transport in wild-type cells, residual nucleoside transport in AE1 cells was resistant to inhibition by nitrobenzylthioinosine. In its properties the latter resembled nitrobenzylthioinosine-resistant nucleoside transport observed in other types of mammalian cells. It amounted to less than 1% of the total nucleoside transport activity of wild-type S49 cells. The results indicate that nitrobenzylthioinosine-resistant and -sensitive nucleoside transports are genetically distinguishable. In wild-type cells, the salvage of thymidine, when present at concentrations higher than 1 to 10 microM, was limited by phosphorylation, because of the saturation of thymidine kinase. In AE1 cells, entry into the cells mainly limited thymidine salvage, but at high thymidine concentrations the combined entry via residual transport and nonmediated permeation was sufficiently rapid to support intracellular thymidine phosphorylation at rates comparable to those observed in wild-type cells.  相似文献   

2.
Summary This study presents the first evidence that the 5 promoter region of the Saccharomyces cerevisiae glyceraldehyde-3-phosphate dehydrogenase gene (G-3-PD) promoter will permit expression of an adjacent foreign gene. The S. cerevisiae G-3-PD promoter was linked to the herpes simplex virus — thymidine kinase (HSV-TK) gene in a shuttle plasmid capable of autonomous replication in both yeast and Escherichia coli. Since the HSV-TK gene promoter is not functional in yeast, yeast cells containing these plasmids will express the HSV-TK gene and synthesize thymidine kinase only if the yeast promoter fragment is fused to the HSV-TK gene in the proper orientation. The 5 flanking sequences necessary for the expression of heterologous eukaryotic genes in S. cerevisiae are discussed.  相似文献   

3.
The relationship between changes in the intracellular free Ca2+ concentration, [Ca2+]i, and the initiation of proliferation of murine B cells after the addition of mitogens and activators was studied. The effects of lipopolysaccharide (LPS), 12-O-tetradecanoyl phorbol-13-acetate (TPA), rabbit IgG antimouse Fab (IgG RAM Fab), and its F(ab')2 fragment (F(ab')2 anti-Fab) on the [Ca2+]i were measured using the fluorescent calcium indicator Fura-2. In parallel experiments, DNA and/or RNA synthesis were measured by assaying [3H]thymidine and/or [3H]uridine uptake. LPS stimulated a 20-120 X increase in the [3H]thymidine uptake, and a 3-7 X increase in [3H]uridine uptake without inducing any change in the [Ca2+]i. TPA induced a marginal increase in [3H]thymidine and [3H]uridine uptake, without effecting any change in the [Ca2+]i. In contrast, low doses of IgG RAM Fab produced a triphasic change in the [Ca2+]i, but had no effect on the [3H]thymidine or [3H]uridine uptake, even at much higher concentrations. Similarly, low doses of the F(ab')2 fragment induced sizable increases in the [Ca2+]i without affecting the [3H]nucleoside uptake. However, higher concentrations of F(ab')2 anti Fab increased the [3H]thymidine uptake and [3H]uridine uptake, while also increasing the [Ca2+]i. Significantly, pretreating the cells with TPA for 3 min virtually abolished the [Ca2+]i increase induced by IgG RAM Fab while simultaneously potentiating an increase in the IgG RAM Fab-induced [3H]thymidine uptake 85-fold. In the presence of TPA, IgG RAM Fab also induced a 2- to 30-fold increase in [3H]uridine uptake. Similarly, TPA virtually abolished the [Ca2+]i increase induced by the F(ab')2 anti-Fab fragment, yet it stimulated a F(ab')2 anti-Fab-induced uptake of [3H]thymidine and [3H]uridine by 120 and 10 times, respectively.  相似文献   

4.
The structural changes taking place in the enzyme thymidine phosphorylase (TPase, also known as PD-ECGF) that are required to achieve catalytic competence upon binding thymidine and phosphate have been simulated by means of targeted molecular dynamics (tMD). The hinge regions were characterized by structural homology comparisons with pyrimidine nucleoside phosphorylase, whose X-ray structure has been solved both in a closed and in an open form. The rearrangement of residues around the substrate that was observed during the tMD trajectory suggested that His-85 could be playing an important role in the catalytic mechanism. A quantum mechanical study of the reaction in the presence of the most relevant active site residues was then performed at the semiempirical level. The results revealed that His-85 could be involved in the protonation of the pyrimidine base at the O2 position to yield the enol tautomer of the base. To establish the role of this oxygen atom in the reaction, ground states, transition states, and final products were studied using higher level ab initio methods starting from both thymidine and 2-thiothymidine as alternative substrates. Comparison of both transition states showed that replacing the oxygen at position 2 of the pyrimidine base by sulfur should accelerate the reaction rate. Consistent with this result, 2-thiothymidine was shown to be a better substrate for TPase than the natural substrate, thymidine. For simulating the final step of the reaction, tMD simulations were used to study domain opening upon product formation considering both the enol and keto tautomers of thymine. Product release from the enzyme was easiest in the simulation that incorporated the keto tautomer of thymine, suggesting that the enol intermediate spontaneously tautomerizes back to the more energetically stable keto form. These results highlight a previously unreported role for His-85 in the catalytic mechanism of TPase and can have important implications for the design of novel TPase inhibitors.  相似文献   

5.
The Daudi line of human lymphoblastoid cells shows a high sensitivity towards growth inhibition by human interferons. In cells pretreated with 70 reference units/ml of an interferon preparation for 48 h, the incorporation of exogenous [3H]thymidine into DNA is inhibited by as much as 85%. We are investigating the extent to which this effect reflects a true inhibition of the rate of DNA synthesis or whether it may be caused by changes in the metabolic utilization of exogenous thymidine by the cells. Interferon treatment results in a 30% inhibition of the rate of membrane transport and a 60% decrease in the rate of phosphorylation of [3H]thymidine in vivo. The latter effect is due to a decrease in V of thymidine kinase without any change in the value of Km for this enzyme. In addition to these changes, incorporation of [3H]uridine into DNA, which occurs as a result of the intracellular conversion of this precursor into thymidine nucleotides, is also inhibited by 75%, whereas RNA labelling by [3H]uridine is decreased by only 15% in interferon-treated cells. Thus several different metabolic events associated with thymidine nucleotide metabolism and DNA synthesis in Daudi cells are disrupted by interferon treatment.  相似文献   

6.
Abstract: The localization and mechanism of thymidine and deoxyuridine transport in the central nervous system were studied in vivo and in vitro . Previous studies have shown that thymidine enters brain from blood in part via the CSF. In vitro , isolated adult bovine cerebral microvessels, which readily concentrated and phosphorylated deoxyglucose, were unable to concentrate thymidine and deoxyuridine. In vivo , [3H]thymidine (0.2 μ M ) and [3H]deoxyuridine(0.4 μ M ) were not extracted more readily than [14C]sucrose in a single pass through the cerebral circulation of rats. In vivo , [3H]thyrnidine retention in CSF and brain after entry from blood was increased when the efflux of [3H]thymidine from CSF and the phosphorylation of [3H]thymidine in brain were depressed by the intraventricular injection of unlabeled thymidine. These studies and previous work suggest that the transfer of thymidine (and deoxyuridine) through the blood-brain barrier in either direction must be extremely low. The present studies are consistent with the postulate that thymidine is transported by an active transport system in the choroid plexus that transfers thymidine from blood into the CSF; from the CSF, the thymidine enters brain cells and is phosphorylated.  相似文献   

7.
The murine V beta 2 promoter was analyzed for an element regulating phorbol ester inducibility of the TCR beta chain gene. In transient expression analysis of 5' nested deleted fragments of the V beta 2 promoter, the TPA-inducible element mapped between -85 and -42. The -85 to -62 oligo conferred 12-0-tetradecanoylphorbol-13-acetate (TPA) inducibility to the heterologous TPA-uninducible thymidine kinase promoter. The -85 to -62 region contained an AP-1 site (-85 to -72) and inverted repeat motif (-72 to -62). The AP-1 site required the 3' flanking inverted repeat region for conferring optimal inducibility. In vitro transcribed and translated jun/fos heterodimers bind to the V beta 2 AP-1 motif with a 16-fold lower affinity as compared to the collagenase AP-1 motif. This explains the inability of the V beta 2 AP-1 motif to confer optimal TPA inducibility by itself. The affinity of jun/fos heterodimers for the V beta 2 AP-1 motif was not increased by the presence in cis of the inverted repeat motif. The 3' flanking inverted repeat binds the ets transactivator but not jun/fos heterodimers. The demonstrated cooperativity between the AP-1 and the 3' flanking sequence to confer TPA inducibility can thus be explained by the individual contributions of jun/fos and ets transactivators.  相似文献   

8.
Techniques are described by which the transport of nutrients into mammalian cells in suspension can be measured at intervals of 1.5 seconds. By application of these techniques, the existence of a saturable (Km = 85 muM), non-concentrative, transport system for thymidine was demonstrated in Novikoff rat hepatoma cells depleted of ATP. At concentrations of thymidine less than the Km, this system operated at velocities sufficient to nearly completely equilibrate intra- and extra-cellular thymidine pools within 8 seconds. In phosphorylating cells, the transport system operated with similar rapidity, so that intracellular phosphorylation was rate-limiting for the incorporation of thymidine into nucleotides. Uptake of 3-O-methylglucose occurred at comparable velocities, attaining 90% of equilibrium between internal and external pools within 25 seconds. Uptake of cytosine by simple diffusion was 100 times slower.  相似文献   

9.
Thymidine incorporation in nucleoside transport-deficient lymphoma cells   总被引:4,自引:0,他引:4  
Nucleoside transport deficiency in mammalian cells is associated with an inability to transport most nucleosides, growth resistance to a spectrum of cytotoxic nucleosides, and a loss of binding sites for 4-nitrobenzylthioinosine (NBMPR), a potent inhibitor of nucleoside transport. The nucleoside transport-deficient S49 T lymphoma cell line, AE1, however, was almost as capable of incorporating thymidine into TTP as the wild type parent provided thymidine was administered at a sufficiently high concentration. Consequently, AE1 cells were just as sensitive as wild type cells to the toxicity of high thymidine concentrations. In contrast, AE1 cells were highly resistant to almost all other cytotoxic nucleosides including the thymidine analogs, 5-bromodeoxyuridine and 5-fluoro-2'-deoxyuridine 5'-monophosphate. Despite having demonstrable ability to accumulate TTP, AE1 cells were unable to grow on hypoxanthine-amethopterin-thymidine (HAT)-containing medium. This was due to their inability to accumulate sufficient TTP from the low concentrations of thymidine present in HAT medium. AE1 cells possessed an incomplete thymidine transport deficiency, the extent of which was concentration dependent. The residual capacity for thymidine transport present in AE1 cells was insensitive to inhibition by 4-nitrobenzylthioinosine and could account both for their inability to grow on HAT medium and their sensitivity to cytotoxic concentrations of thymidine. Another nucleoside transport-deficient cell line, FURD-80-3-6, was similar to the AE1 cell line in its growth phenotype and NBMPR-binding site deficiency but differed in its decreased growth sensitivity to thymidine. That nucleoside transport deficiencies may vary in their completeness for different nucleosides has significance for the mechanism by which a single transporter can recognize a wide variety of nucleosides.  相似文献   

10.
The kinetics of thymidine uptake by Escherichia coli and Bacillus subtilis cells in the presence of adenine and guanine nucleosides was investigated. The initial concentration of thymidine in the growth medium was 0.35 microng/ml while the initial concentration of purine nucleosides ranged from 25 to 250 microng/ml. Adenine nucleosides when present at a concentration more than 50 microng/ml strongly inhibit thymidine uptake by the bacteria. The duration of the inhibition depends on the initial concentration of adenine nucleoside in the growth medium. At an initial concentration of deoxyadenosine (or adenosine) of 250 microng/ml the time of inhibition of thymidine uptake was about 60 min. During this period thymidine is almost completely preserved from the action of bacterial thymidine phosphorylase. Guanine nucleosides (guanosine or deoxyguanosine) do not markedly inhibit thymidine uptake by bacteria even at a concentration of 250 microng/ml. It is shown that they do protect thymidine from the phosphorolytic action of the thymidine phosphorylase although much less effectively than adenine nucleosides. It is suggested that some areas in the bacterial membrane where thymidine phosphorylase is located are not available to guanine nucleosides.  相似文献   

11.
A comparison was made between the utilization of thymine and thymidine for the synthesis of DNA in Novikoff hepatoma cells growing in suspension culture. When the cell cultures were switched from exponential growth to a relatively non-growing condition, by resuspending them in culture media minus serum for 18 h, there was an 85% decrease in the rate of thymidine incorporation but only a 15% decrease in the rate of thymine incorporation. Exposure to an alkylating agent (methyl methane sulfonate) resulted in a 79% decrease in thymidine incorporation, while thymine incorporation was decreased only 35%. Thymidine at a concentration equal to its Km for incorporation into DNA (4 × 10−7 M) had virtually no effect on thymine incorporation. It was not until a thymidine concentration of ten times the Km was employed that appreciable (40%) decreases in the rate of thymine incorporation were observed. Examination of total cellular DNA or nuclear DNA gave similar results. These studies are interpreted as indicating the presence of multiple precursor pools for the synthesis of DNA-thymine in Novikoff hepatoma cells.  相似文献   

12.
Incorporation of thymidine into Novikoff rat hepatoma cells was analyzed with a rapid sampling technique which allowed collection of 12 time points in 20 sec. Transport was studied in the absence of metabolism by using either ATP-depleted cells or a thymidine kinase negative subline. Transport was a rapid, saturable, nonconcentrative process with a Km of about 85 μM. The intracellular thymidine pool was also rapidly labeled in cells which phosphorylated thymidine, so that a group translocation process involving thymidine kinase can be ruled out. Under all conditions examined, phosphorylation, not the transport, of thymidine was the rate-determining step in its incorporation into the acid-soluble pool. Estimation of transport rates from total incorporation into cells which phosphorylate the substrate is invalid in this cell system and must be questioned in all instances.  相似文献   

13.
In our previous study, bradykinin (BK) exerts its mitogenic effect through Ras/Raf/MEK/MAPK pathway in vascular smooth muscle cells (VSMCs). In addition to this pathway, the non-receptor tyrosine kinases (Src), EGF receptor (EGFR), and phosphatidylinositol 3-kinase (PI3-K) have been implicated in linking a variety of G-protein coupled receptors to MAPK cascades. Here, we investigated whether these different mechanisms participating in BK-induced activation of p42/p44 MAPK and cell proliferation in VSMCs. We initially observed that BK- and EGF-dependent activation of Src, EGFR, Akt, and p42/p44 MAPK and [3H]thymidine incorporation were mediated by Src and EGFR, because the Src inhibitor PP1 and EGFR kinase inhibitor AG1478 abrogated BK- and EGF-dependent effects. Inhibition of PI3-K by LY294002 attenuated BK-induced Akt and p42/p44 MAPK phosphorylation and [3H]thymidine incorporation, but had no effect on EGFR phosphorylation, suggesting that EGFR may be an upstream component of PI3-K/Akt and MAPK in these responses. This hypothesis was supported by the tranfection with dominant negative plasmids of p85 and Akt which significantly attenuated BK-induced Akt and p42/p44 MAPK phosphorylation. Pretreatment with U0126 (a MEK1/2 inhibitor) attenuated the p42/p44 MAPK phosphorylation and [3H]thymidine incorporation stimulated by BK, but had no effect on Akt activation. Moreover, BK-induced transactivation of EGFR and cell proliferation was blocked by matrix metalloproteinase inhibitor GM6001. These results suggest that, in VSMCs, the mechanism of BK-stimulated activation of p42/p44 MAPK and cell proliferation was mediated, at least in part, through activation of Src family kinases, EGFR transactivation, and PI3-K/Akt.  相似文献   

14.
Class IA phosphoinositide 3-kinase (PI3K) is involved in regulating many cellular functions including cell growth, proliferation, cell survival, and differentiation. The p85 regulatory subunit is a critical component of the PI3K signaling pathway. Mesenchymal stem cells (MSC) are multipotent cells that can be differentiated into osteoblasts (OBs), adipocytes, and chondrocytes under defined culture conditions. To determine whether p85α subunit of PI3K affects biological functions of MSCs, bone marrow-derived wild type (WT) and p85α-deficient (p85α(-/-)) cells were employed in this study. Increased cell growth, higher proliferation rate and reduced number of senescent cells were observed in MSCs lacking p85α compare with WT MSCs as evaluated by CFU-F assay, thymidine incorporation assay, and β-galactosidase staining, respectively. These functional changes are associated with the increased cell cycle, increased expression of cyclin D, cyclin E, and reduced expression of p16 and p19 in p85α(-/-) MSCs. In addition, a time-dependent reduction in alkaline phosphatase (ALP) activity and osteocalcin mRNA expression was observed in p85α(-/-) MSCs compared with WT MSCs, suggesting impaired osteoblast differentiation due to p85α deficiency in MSCs. The impaired p85α(-/-) osteoblast differentiation was associated with increased activation of Akt and MAPK. Importantly, bone morphogenic protein 2 (BMP2) was able to intensify the differentiation of osteoblasts derived from WT MSCs, whereas this process was significantly impaired as a result of p85α deficiency. Addition of LY294002, a PI3K inhibitor, did not alter the differentiation of osteoblasts in either genotype. However, application of PD98059, a Mek/MAPK inhibitor, significantly enhanced osteoblast differentiation in WT and p85α(-/-) MSCs. These results suggest that p85α plays an essential role in osteoblast differentiation from MSCs by repressing the activation of MAPK pathway.  相似文献   

15.
Two heat-sensitive (reversibly arrested in G1 phase at 39.5 degrees C, multiplying at 33 degrees C) and two cold-sensitive (reversibly arrested in G1 phase at 33 degrees C, multiplying at 39.5 degrees C) cell-cycle mutants of the P-815-X2 murine mastocytoma line were tested for ribonucleotide reductase activity, using cells made permeable to nucleotides. After transfer of the heat-sensitive mutant cells to 39.5 degrees C, ribonucleotide reductase activity, similar to thymidine kinase (Schneider, E., Müller, B. and Schindler, R. (1983) Biochim. Biophys. Acta 741, 77-85), but unlike DNA polymerase alpha (Schneider, E., Müller, B. and Schindler, R. (1985) Biochim. Biophys. Acta 825, 375-383), decreased rapidly and in parallel with numbers of cells in S phase, whereas in the cold-sensitive mutant cells brought to 33 degrees C, ribonucleotide reductase activity decreased approx. 8 h later than numbers of DNA-synthesizing cells. When arrested heat- or cold-sensitive mutant cells were returned to the permissive temperature, ribonucleotide reductase activities, similar to DNA polymerase alpha and to thymidine kinase in heat-sensitive mutants, increased essentially in parallel with reentry of cells into S phase, whereas the increase in thymidine kinase activity in the cold-sensitive mutants was previously shown to occur approx. one cell-cycle time later. This indicates that ribonucleotide reductase and thymidine kinase are coordinately expressed in the heat-sensitive, but independently regulated in the cold-sensitive mutants.  相似文献   

16.
The enzyme thymidine phosphorylase (thymidine: orthophosphate deoxyribosyltransferase, EC 2.4.2.4.), which plays a crucial role in nucleic acid metabolism in both prokaryotic and eukaryotic cells by regulating the availability of thymidine, is present in mammalian blood. Here we describe a simple, rapid HPLC-based micromethod for the assay of blood thymidine phosphorylase. We have arbitarily defined 1 unit of blood thymidine phosphorylase activity as the activity required to produce a 1-nM increment in the plasma concentration of thymine after incubation for 1 h at 37°C with a saturating concentration of exogenous thymidine.

In normal adults, whole (peripheral venous) blood thymidine phosphorylase activity with blood cells intact was 64 ± 11 units (mean ± S.D., n =20, range 45–89). The apparent Michaelis constant for thymidine was of the order to 10−4 M but varied nearly 5-fold between different individuals. Activity increased when blood cells were permeabilised or lysed with non-ionic detergents, implying that thymidine phosphorylase is an intracellular enzyme which may be influenced by exogenous as well as intracellular factors. When blood from normal donors was fractionated, thymidine phosphorylase activity consistently co-isolated with platelets. Whole-blood thymidine phosphorylase activity correlated well with platelet parameters. Although thymidine phosphorylase activity was also detected in plasma and serum, the small size and notorious fragility of platelets suggest its platelet origin.

Blood from leukaemic donors showed significantly increased thymidine phosphorylase activity compared to normal controls (mean activity ± S.D. was 96 ± 27 units; range 58–140, n = 8).

Thymine formation from thymidine was temperature- and pH-depdendent in whole blood. 2′-Deoxyuridine and 3 of its 5-halogenated analogues (but not 3′-azido-3′-deoxythymidine (AZT), were catabolised by blood thymidine phosphorylase, even during blood clotting at room temperature. Assumptions about in vivo concentrations of these compounds should therefore be interpreted cautiously.

In the presence of high concentrations of thymine and suitable deoxyribose donors, small amounts of thymidine were formed in some blood samples, so it is conceivable that thymidine catabolism may be reversible in vivo under some circumstances.  相似文献   


17.
The partially self-complementary synthetic DNA oligonucleotide d(CG)5T4(CG)5 has been studied by using 1H and 31P NMR and circular dichroism. Results show that, under low-salt conditions (120 mM NaCl buffer), an intramolecular hairpin loop exists in which the double-helical stem region is B-form and the thymidine loop residues have predominantly southern (C2'-endo) sugar conformations. The thymidine glycosidic torsion angles are intermediate between syn and anti or exist as an equilibrium mixture of residues in the two extremes. NOESY data indicate that the structure of the loop region is very similar to that found for d(CG)2T4(CG)2 [Hare, D. R., & Reid, B. R. (1986) Biochemistry 25, 5341-5350]. Under high-salt conditions (6 M NaClO4 buffer), the dominant form (approximately equal to 85%) is an intramolecular hairpin structure in which the stem region forms a Z-form double helix. As in the B-form, the loop thymidine residues are intermediate between the syn and anti conformations or exist as an equilibrium mixture of the two, but the thymidine sugar conformations differ in that they are biased toward northern (C3'-endo) conformations.  相似文献   

18.
A mutant of Escherichia coli lacking the catabolic enzyme thymidine phosphorylase readily incorporates exogenous thymidine into deoxyribonucleic acid (DNA) even when provided at concentrations as low as 0.2 mug/ml. Incorporation by this prototrophic strain occurs specifically into DNA, since, with radioactively labeled thymidine, (i) more than 98% is incorporated into alkali-stable material, (ii) at least 90% is recovered as thymine after brief formic acid hydrolysis, and (iii) at least 90% is incorporated into material with the buoyant density of DNA. During growth in medium containing thymidine, the bacteria obtain approximately half of their DNA thymines from the exogenous thymidine and half from endogenous synthesis. The thymines and cytosines of DNA can be simultaneously and specifically labeled by thymidine-2-(14)C and uridine-5-(3)H, respectively. The mutant, which does not degrade thymidine, retains the ability to degrade the thymidine analogue 5-bromodeoxyuridine.  相似文献   

19.
Abstract: The bacterial community response to pH was studied for 16 soils with pH(H2O) ranging between 4 and 8 by measuring thymidine incorporation into bacteria extracted from the soil into a solution using homogenization-centrifugation. The pH of the bacterial solution was altered to six different values with dilute sulfuric acid or different buffers before measuring incorporation. The resulting pH response curve for thymidine incorporation was used to compare bacterial communities from the different soils. There was a correlation between optimum pH for thymidine incorporation and the soil pH(H2O). Even bacterial communities from acid soils had optima corresponding to the soil pH, indicating that they were adapted to these conditions. Thymidine incorporation was also compared with leucine incorporation for some soils. The leucine to thymidine incorporation ratio was constant over the tested pH interval when incorporation values were adjusted for isotope dilution. A good correlation was found between the scores along the first component (explaining 80% of the variation) and soil pH ( r 2 = 0.85), if principal component analysis of the pH response curves for thymidine incorporation was used. The pH response curves differed most for the extreme pH values used, and a linear relationship was found between the logarithm of the ratio of thymidine incorporation at pH 4.3 to incorporation at pH 8.2 and the soil pH ( r 2 = 0.86). Thus, a simplified technique using only two pH values, when measuring the thymidine incorporation, could be used to compare the response to pH of bacterial communities.  相似文献   

20.
R J Pinney 《Microbios》1977,19(75):55-66
Minimum satisfactory concentrations of thymine and thymidine were determined for the growth of a high thymine-requirng (thy) mutant to Escherichia coli strain J5-3. Cultures were then grown in the presence of these concentrations of non-radioactive ('cold') pyrimidine together with 5 microCi/ml [methyl-3H)thymine, or [methyl-3H)thymidine (specific activities 5 Ci/m mole), and the uptake of radioactivity into ice cold trichloroacetic acid insoluble material determined. By far the most efficient labelling system was obtained if the label was supplied as radioactive thymidine and growth requirements satisfied by thymine alone. The addition of deoxyadenosine to the labelled thymidine/unlabelled thymine system dramatically reduced uptake of label. The addition of radioactive thymine with either thymine or thymidine to ensure satisfactory growth gave poor labelling. Using the [methyl-3H] thymidine/thymine system it was possible to increase the concentration of thymine from 8 to 64 microgram/ml with only a 25% reduction in label uptake after a 2 h period. The same system was also shown to be most efficient for labelling a thy derivative of another K12 strain, a thymine low-requiring (tir) K12 strain, a thy mutant of Klebsiella aerogenes 418 and a tir derivative of Salmonella typhimurium LT2.  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