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 20 毫秒
1.
We used a direct polymerase chain reaction (PCR) method for quantification of HPRT exons 2+3 deletions and t(14;18) translocations as a measure of illegitimate V(D)J recombination. We determined the baseline frequencies of these two mutations in mononuclear leukocyte DNA from the umbilical cord blood of newborns and from the peripheral blood of adults. In an initial group of 21 newborns, no t(14;18) translocations were detected (<0.049×10−7). The frequency of HPRT exons 2+3 deletions was 0.10×10−7 per mononuclear leukocyte, lower than expected based on the T-cell proportion of this cell fraction (55%–70%) and previous results using the T-cell cloning assay (2–3×10−7 per clonable T-cell). Phytohemagglutinin (PHA), as used in the T-cell cloning assay, was examined for its effect on the frequencies of these mutation events in mononuclear leukocytes from an additional 11 newborns and from 12 adults. There was no significant effect of PHA on t(14;18) translocations which were rare among the newborns (1 detected among 2.7×108 leukocytes analyzed), and which occurred at frequencies from <1×10−7 (undetected) to 1.6×10−4 among the adults. The extremely high frequencies of t(14;18)-bearing cells in three adults were due mainly to in vivo expansion of two to six clones. However, PHA appeared to stimulate a modest (although not significant) increase in the frequency of HPRT exons 2+3 deletions in the leukocytes of the newborns, from 0.07×10−7 to 0.23×10−7. We show that both the direct PCR assay and the T-cell cloning assay detect similar frequencies of HPRT exons 2+3 deletions when calculations are normalized to blood volume, indicating that the apparent discrepancy is probably due to the different population of cells used in the assays. This direct PCR assay may have utility in characterizing the effects of environmental genotoxic agents on this clinically important recombination mechanism.  相似文献   

2.
A pilot biomarker study was conducted to investigate the feasibility of using the hypoxanthine guanine phosphoribosyltransferase (HPRT) gene in peripheral blood lymphocytes as a biomarker for detecting genetic effects of arsenic exposure. Blood and urine samples were obtained from workers highly exposed to arsenic in a copper roasting plant in Antofagasta, Chile. Individuals were classified according to their job titles into three potential exposure groups: high, medium, and low. To confirm exposure, arsenic concentration was determined in urine samples. The HPRT mutant frequencies were measured in lymphocytes from 15 individuals ranging in age from 24 to 66 years. The mean mutant frequencies for the three exposure groups were: low (9×10−6), medium (11×10−6), and high (24×10−6). An increased mutant frequency was observed in the highly exposed group, but the response was so slight that it is not likely that this assay will be capable of providing dose–response information across a range of lower, more typical environmental arsenic levels.  相似文献   

3.
We investigated the effect of NDMA and DNSGU on the induction of chromosomal aberrations and sister-chromatid exchanges (SCEs), as well as the influence of the former compound on cell-cycle kinetics in cultured cow peripheral lymphocytes. A clastogenic effect was observed in treated cell cultures at 6 or 12 × 10−5 M concentrations of NDMA and DNSGU, respectively, but no increase of chromosomal breaks was seen at the lowest dose. NDMA at 6 × 10−4 M was toxic to cow lymphocytes. NDMA and DNSGU induced statistical increases of SCEs at the test doses (6 or 12 × 10−6 and 6 or 12 × 10−5 M, respectively). In addition, treatment with NDMA at a dose of 6 × 10−5 M revealed significant heterogeneity of the first, second and third metaphases between treated and untreated groups. A reduction of the proliferation index and proliferation delay per cycle was shown too.  相似文献   

4.
Palythoa psammophilia Walsh & Bowers has a well coordinated, stereotyped feeding response, the culminating step of which is ingestion; this may be elicited by the synergistic effect of the tripeptide glutathione and the -imino acid, proline. Either activator acting separately causes responses only at high concentrations (above 10−5 M for glutathione; above 10−4 M for proline) in a reduced number of animals and at a low rate (5.00 ± 1.73 min in 5 × 10−3 M solutions of glutathione; 11.10±3.74 min in 5 × 10−3 M solutions of proline). Highest percentages of response were obtained in combinations where glutathione was at a concentration of 5 × 10−3 M and proline at 5 × 10−4 M or in combinations of glutathione at concentrations 5 × 10−6 M and proline at 5 × 10−5 M. The speed of ingestion is considerably enhanced when these activators are combined (1.17±1.18 min).  相似文献   

5.
Guar gum has been modified by graft copolymerization with acrylic acid in aqueous medium using vanadium (V)–mercaptosuccinic acid redox system. The optimum reaction conditions affording maximum grafting ratio, efficiency, add on and conversion have been determined. The grafting parameters have been found to increase with increase in vanadium (V) concentration upto 1.0 × 10−2 mol dm−3, but these parameters decrease on further increasing the vanadium (V) concentration. On increasing the mercaptosuccinic acid concentration from 1.0 × 10−2 to 4.0 × 10−2 mol dm−3 grafting ratio, efficiency and add on increase up to 2.0 × 10−2 mol dm−3 but decrease with further increase in mercaptosuccinic acid concentration. On varying the acrylic acid concentration from 5.0 × 10−2 to 30.0 × 10−2 mol dm−3, maximum grafting ratio, efficiency and add on have been obtained at 20.0 × 10−2 mol dm−3. The grafting ratio, add on and conversion increase, on increasing the H+ ion concentration from 1.5 × 10−1 to 6.0 × 10−1 mol dm−3. On increasing the guar gum concentration the grafting parameters increase. The grafting ratio, add on and conversion have been found to increase with time period while efficiency started decreasing after 120 min. It has been observed that %G increases on increasing the temperature up to 35 °C. The graft copolymer has been characterized by IR spectroscopy and thermogravimetric analysis.  相似文献   

6.
Graft copolymer of k-carrageenan and N,N-dimethylacrylamide has been synthesized by free radical polymerization using peroxymonosulphate/glycolic acid redox pair in an inert atmosphere. The grafting parameters i.e. grafting ratio, add on and efficiency decrease with increase in concentration of k-carrageenan from 0.6 to 1.4 g dm−3 and hydrogen ion from 3 × 10−3 to 7 × 10−3 mol dm−3, but these grafting parameters increase with increase in concentration of N,N-dimethylacrylamide from 16 × 10−2 to 32 × 10−2 mol dm−3, and peroxymonosulphate from 0.8 × 10−2 to 2.4 × 10−2 mol dm−3. The metal ion sorption, swelling behaviour and flocculation properties have been studied. The intrinsic viscosity of pure and grafted samples has been measured by using Ubbelohde capillary viscometer. Flocculation capability of k-carrageenan and k-carrageenan-g-N,N-dimethylacrylamide for both coking and non-coking coals has been studied for the treatment of coal mine waste water. The graft copolymer has been characterized by Infrared (IR) spectroscopy and thermogravimetric analysis.  相似文献   

7.
We have designed and synthesized a series of small peptides containing a perfluoroalkyl ketone group at the C-terminal position of the angiotensin I sequence as inhibitors of human renin. From this series of compounds, 8 and 10 showed strong inhibition of human renin (IC50 = 3 × 10−9, 7 × 10−9 M, respectively). Compound 10 did not inhibit pepsin and cathepsin D at 10−4 M. Comparison of the IC50 of compound 8 and compound 11 (8.7 × 10−7 M) demonstrated the marked effect of the perfluoropropyl group on the potency of inhibition on renin, presumably due to the strong electron-withdrawing effect causing the ketone in 8 to exist predominantly as the hydrate — thus mimicking the tetrahedral transition state during hydrolysis of the scissile Leu10—Val11 amide bond.  相似文献   

8.
Both prostaglandins (PGs) and nitric oxide (NO) have cytoprotective and hyperemic effects in the stomach. However, the effect of NO on PG synthesis in gastric mucosal cells is unclear. We examined whether sodium nitroprusside (SNP), a releaser of NO, stimulates PG synthesis in cultured rabbit gastric mucus-producing cells. These cells did not release NO themselves. Co-incubation with SNP (2 × 10−4, 5 × 10−4, 10−3 M) increased PGE2 synthesis, and SNP (10−3 M) increased PGI2 synthesis in these cells. Hemoglobin, a scavenger of NO, (10−5 M) eliminated the increase in PGE2 synthesis by SNP, but methylene blue, an inhibitor of soluble guanylate cyclase, (5 × 10−5 M) did not affect the increase in PGE2 synthesis by SNP. 8-bromo guanosine 3′ : 5′-cyclic monophosphate (8-bromo cGMP), a cGMP analogue, (10−6, 10−5, 10−4, 10−3 M) did not affect PGE2 synthesis. These findings suggest that NO increased PGE2 and PGI2 synthesis via a cGMP-independent pathway in cultured rabbit gastric cells.  相似文献   

9.
The authors incubated adrenal mitochondria to study the in vitro action of cortisol and testosterone on the transformation of corticosterone and 18-hydroxycorticosterone into aldosterone. The results show that cortisol at concentrations of 5 × 10−6 and 10−4 M inhibit the conversion of corticosterone into aldosterone by 23.6 to 90%; testosterone 5 × 10−5 and 10−4 M inhibit the reaction by 78.4 and 87.2%, respectively. The inhibition of the conversion of 18-hydroxycorticosterone into aldosterone is 12.5 to 91% by cortisol with concentrations ranging from 5 × 10−7 to 5 × 10−5 M and testosterone 5 × 10−5 and 10−4 M inhibits the reaction by 87.3 and 91%, respectively. Aldosterone (10−8 and 10−6 M) does not inhibit aldosterone biosynthesis from corticosterone or 18-hydroxycorticosterone. It thus appears that cortisol and testosterone have an effect on the aldosterone biosynthesis pathways in mitochondria. This action may be located at the binding site of the cytochrome P450 11β, which catalyzes all hydroxylation steps in the mineralocorticoid biosynthesis pathway. Because cortisol and testosterone may interfere with aldosterone biosynthesis, and since functional zonation is expected in adrenal carcinomas, the presence of these steroids in substantial amounts could explain the very low plasma aldosterone level usually observed, in adrenal carcinomas studies in our laboratory.  相似文献   

10.
Tang ML  Wang SC  Wang T  Zhao SG  Wu YJ  Wu LJ  Yu ZL 《Mutation research》2006,602(1-2):163-169
The mutational spectrum of the genomic lacI gene induced by low-energy nitrogen ion irradiation in wild type Escherichia coli strain W3110 were compared with the spontaneous and the vacuum controls. The mutant frequency of irradiated group was dose-dependent and reached 26.3 × 10−6 at dose of 31.2 × 1014 ions/cm2, which was about 18-fold over the background (1.5 × 10−6) and 10-fold over the vacuum controls (2.6 × 10−6). This result indicated that the low-energy ion irradiation was one of many effective mutagens, though the vacuum condition of low-energy ions contributed some low-level gene mutations. It was found that the difference between the spontaneous and the vacuum control was the increases of base-pair substitutions in the vacuum control group. The spectra of irradiated group were quite similar to that of oxygen free-radical induced in the same strain, suggesting free-radicals and other adducts generated by low-energy ions might play an important role in the mutagenesis in vivo. When the spontaneous and the vacuum control group were compared, base-pair substitutions, deletions and additions of the irradiated group were significantly increased, and the +TGGC or −TGGC at hot spot was decreased from 82 to 48%. But the remarkable increase in absolute MF of the +TGGC or −TGGC at hot spot in the irradiated group suggested that low-energy ions did induce the mutations of this type. The spectra of our irradiated group had relative low-level base-pair substitutions, high-level ±TGGC and high proportion additions than those of γ-radiation induced, implying there were some different effects or processes between them.  相似文献   

11.
It is well recognized that estradiol (E2) is one of the most important hormones supporting the growth and evolution of breast cancer. Consequently, to block this hormone before it enters the cancer cell or in the cell itself, has been one of the main targets in recent years. In the present study we explored the effect of the progestin, nomegestrol acetate, on the estrone sulfatase and 17β-hydroxy-steroid dehydrogenase (17β-HSD) activities of MCF-7 and T-47D human breast cancer cells. Using physiological doses of estrone sulfate (E1S: 5 × 10−9 M), nomegestrol acetate blocked very significantly the conversion of E1S to E2. In the MCF-7 cells, using concentrations of 5 × 10−6 M and 5 × 10−5 M of nomegestrol acetate, the decrease of E1S to E2 was, respectively, −43% and −77%. The values were, respectively, −60% and −71% for the T-47D cells. Using E1S at 2 × 10−6 M and nomegestrol acetate at 10−5 M, a direct inhibitory effect on the enzyme of −36% and −18% was obtained with the cell homogenate of the MCF-7 and T-47D cells, respectively. In another series of studies, it was observed that after 24 h incubation of a physiological concentration of estrone (E1: 5 × 10−9 M) this estrogen is converted in a great proportion to E2. Nomegestrol acetate inhibits this transformation by −35% and −85% at 5 × 10−7 M and 5 × 10−5 M, respectively in T-47D cells; whereas in the MCF-7 cells the inhibitory effect is only significant, −48%, at 5 × 10−5 M concentration of nomegestrol acetate. It is concluded that nomegestrol acetate in the hormone-dependent MCF-7 and T-47D breast cancer cells significantly inhibits the estrone sulfatase and 17β-HSD activities which converts E1S to the biologically active estrogen estradiol. This inhibition provoked by this progestin on the enzymes involved in the biosynthesis of E2 can open new clinical possibilities in breast cancer therapy.  相似文献   

12.
Data are reported for the binding of Ni2+, Co2+, and Mg2+ to the B-form of double-stranded poly(dG-dC) at ionic strength conditions I = 0.001 M, 0.01 M, and 0.1 M. The apparent binding constants for Ni2+ and Co2+ are about the same and are 2- to 3-fold higher than those for Mg2+. Kinetic studies indicate that Mg2+ binds to the polynucleotide mainly (or solely) as a mobile cloud (electrostatically, outer-sphere), whereas the transition metal ions undergo site binding (inner-sphere coordination) with poly(dG-dC). The kinetic data suggest that an Ni2+ ion coordinates to more than one binding site at the polynucleotide, presumably to G-N7 and a phosphate group.

At low ionic strength conditions the addition of Ni2+ induces a B → Z conformational transition in poly(dG-dC). As demonstrated by UV absorption and CD spectroscopy, the transition occurs at I = 0.001 M already when 3 × 10−5 – 7 × 10−5 M of Ni2+ are added to 8 × 10−5 M (in monomeric units) of poly(dG-dC), and at I = 0.01 M between 2.5 × 10−4 and 4.5 × 10−4 M of Ni2+. Using murexide as an indicator of the concentration of free Ni2+ ions, the amount of Ni2+ which is bound to the polynucleotide could be determined. At I = 0.001 M it was established that the B → Z transition begins when 1 Ni2+ is bound coordinatively per four base pairs, and the transition is complete when 1 Ni2+ is bound coordinatively per three base pairs. It is this coordinated Ni2+ which induces the B → Z transition.  相似文献   


13.
The present work is a study of oxidative degradation of the organic matter present in the washing waters from the black table olive industry. Pollutant organic matter reduction was studied by an aerobic biological process and by the combination of two successive steps: ozonation pretreatment followed by aerobic biological degradation. In the single aerobic biological process, the evolution of biomass and organic matter contents was followed during each experiment. Contaminant removal was followed by means of global parameters directly related to the concentration of organic compounds in those effluents: chemical oxygen demand (COD) and total phenolic content (TP). A kinetic study was performed using the Contois model, which applied to the experimental data, provides the specific kinetic parameters of this model: 4.81×10−2 h−1 for the kinetic substrate removal rate constant, 0.279 g VSS g COD−1 for the cellular yield coefficient and 1.92×10−2 h−1 for the kinetic constant for endogenous metabolism. In the combined process, an ozonation pretreatment is conducted with experiments where an important reduction in the phenolic compounds is achieved. The kinetic parameters of the following aerobic degradation stage are also evaluated, being 5.42×10−2 h−1 for the kinetic substrate removal rate constant, 0.280 g VSS g COD−1 for the cellular yield coefficient and 9.1×10−3 h−1 for the kinetic constant for the endogenous metabolism.  相似文献   

14.
Amperometric choline biosensors were fabricated by the covalent immobilization of an enzyme of choline oxidase (ChO) and a bi-enzyme of ChO/horseradish peroxidase (ChO/HRP) onto poly-5,2′:5′,2″-terthiophene-3′-carboxylic acid (poly-TTCA) modified electrodes (CPMEs). A sensor modified with ChO utilized the oxidation process of enzymatically generated H2O2 in a choline solution at +0.6 V. The other one modified with ChO/HRP utilized the reduction process of H2O2 in a choline solution at −0.2 V. Experimental parameters affecting the sensitivity of sensors, such as pH, applied potential, and temperature were optimized. A performance comparison of two sensors showed that one based on ChO/HRP/CPME had a linear range from 1.0×10−6 to 8.0×10−5 M and the other based on ChO/CPME from 1.0×10−6 to 5.0×10−5 M. The detection limits for choline employing ChO/HRP/CPME and ChO/CPME were determined to be about 1.0×10−7 and 4.0×10−7 M, respectively. The response time of sensors was less than 5 s. Sensors showed good selectivity to interfering species. The long-term storage stability of the sensor based on ChO/HRP/CPME was longer than that based on ChO/CPME.  相似文献   

15.
Biological properties of amino-terminal PTHrP analogues modified in the region 11–13 were examined using ROS 17/2.8 cells. [Leu11,D-Trp12,Arg13,Tyr36]PTHrP(1–36)amide had a 17-fold lower binding affinity for the receptor (apparent Kd: 5 × 10−8 M) than [Tyr36]PTHrP(1–36)amide or [Arg11,13,Tyr36]PTHrP(1–36)amide (apparent Kd for both: 2 × 10−9 M). Moreover, it is only a weak partial agonist despite completely inhibiting radioligand binding. [Leu11,D-Trp12,Arg13,Tyr36,Cys38]PTHrP(7–38) and PTHrP(7–34)amide had similar receptor affinities (apparent Kds: 5 × 10−8 M and 8 × 10−8 M), while that of [Nle8,18,Tyr34]bPTH(7–34)amide was more than 10-fold lower (apparent Kd: 2 × 10−6 M). These changes in biological properties suggest that high affinity receptor binding requires both amino- and carboxyl-terminal domains of the PTHrP(1–36) sequence and/or intramolecular interactions which are impaired by the D-Trp substitution for Gly12.  相似文献   

16.
A new method of physically immobilizing enzymes in poly(2-hydroxyethyl methacrylate) membranes was developed in order to obtain suitable biosensors. It was possible to prepare an enzyme sensor based on an oxygen Clark electrode and on glucose oxidase immobilized by low-temperature gamma radiation-induced polymerization. Temperature and pH effects on the activity of immobilized enzyme are described and the response characteristics of the resulting biosensor are summarized. The determination of glucose in standard solutions was carried out and a linear calibration curve, with an R2 value of 0·9993, from the detection limit 5 × 10−5 to 1·2 × 10−3 was obtained. The biosensor was employed to analysis of control sera and the results were compared to those obtained by enzymatic-spectrophotometric detection.  相似文献   

17.
N-Methyl-N′-nitro-N-nitrosoguanidine (MNNG) reacts with 12 nucleophilic sites in DNA to induce a variety of lesions, but O6-methylguanine (O6-MeG) and O4-methylthymine are the most effective premutagenic lesions produced, mispairing with thymine and guanine, respectively. O6-MeG is repaired by O6-alkylguanine-DNA alkyltransferase (AGT), which removes the methyl group from the O6 position and transfers it to itself, rendering the transferase inactive. When diploid human fibroblasts were exposed to 25 μM, O6-benzylguanine (O6-BzG) in the medium for 3 h, their level of AGT activity was dramatically reduced, to a level of at most 1.6% of the control. Populations of cells pretreated with this level of O6-BzG for 2 h or not pretreated, were exposed to MNNG at a concentration of 2, 4 or 6 μM in the presence or absence of O6-BzG and assayed for survival of colony-forming ability and the frequency of 6-thioguanine-resistant cells (mutations induced in the HPRT gene). O6-BzG (25 μM) was also present in the appropriate half of the cells during the 24 h immediately follwing exposure to MNNG. This 27-h exposure to O6-BzG alone had no cytotoxic or mutagenic effect on the cells but significantly increased the cytotoxicity and mutagenecity of MNNG, increasing the mutant frequency to that found previously in human cells constitutively devoid of AGT activity. At doses of 2 μM and 4 μM MNNG, the mutant frequency observed with the AGT-depleted cells was 120 × 10−6 and 240 × 10−6, respectively; in the cells with abundant AGT activity, these values were 10 × 10−6 and 20 × 10−6, respectively. DNA-sequence analysis of the coding region of the HPRT gene in 36 independent mutants obtained from MNNG-treated AGT-depleted populations and 36 from the control populations showed that even though AGT repair lowered the frequency of mutants by more than 90%, it did not affect the kinds of mutations induced by MNNG nor the strand distribution of the premutagenic guanine lesions. In mutants from the AGT-depleted cells, there were 26 base substitutions and 13 putative splice site mutations; in the control, there were 25 base substitutions and 11 splice site mutations. All but two substitutions involved G · C with 92% being G · C → A · T. In both sets, of the premutagenic lesions were located in the nontranscribed strand. Many ‘hot spots’ were seen, and there was evidence that AGT repaired more lesions from the 5′ half of the gene than from the 3′ half.  相似文献   

18.
The enzymatic activity of mushroom tyrosinase was investigated using catechin as substrate in selected organic solvent media. The results showed that optimal tyrosinase activity was obtained at pH 6.2, 6.6, 6.0 and 6.2 in the organic solvent media of heptane, toluene, dichloromethane, and dichloroethane, respectively, and at a temperature between 25°C and 27.5°C. In addition, the kinetic studies showed that the Km values were 5.38, 1.03, 2.52 and 4.03 mM, for the tyrosinase-catechin biocatalysis in the reaction media of heptane, toluene, dichloromethane, and dichloroethane, respectively, while the corresponding Vmax values were 1.22×10−3, 0.33×10−3, 1.47×10−3 and 1.20×10−3 δA per μg protein per second, respectively. The use of acetone as co-solvent for the tyrosinase-catechin biocatalysis showed that acetone concentrations ranging from 5% to 30% (v/v) in the heptane reaction medium produced a decrease of 4.3% to 96.7% in tyrosinase activity. The results also indicated that the presence of 12.5% acetone in the reaction medium of dichloromethane, and 22.0% in those of toluene and dichloroethane produced a maximal increase of 42.6%, 92.1% and 71.8%, respectively, in tyrosinase activity. However, the overall findings indicated that additional increases in acetone concentration resulted in an inhibition of tyrosinase activity.  相似文献   

19.
R.J.W. De Wit 《FEBS letters》1982,150(2):445-448
Folic acid is degraded too fast by Dictyostelium discoideum to study binding of this ligand to cell surface binding proteins. Folate deaminase activity was inhibited in the presence of 3.3 × 10−4 M 8-azaguanine. This inhibitor enabled us to detect two folate binding proteins. One type bound folic acid and deamino-folic acid with the same affinity (K0.5 = 3–6 × 10−7 M) and apparently negative cooperativity. Binding to only this type was observed if 8-azaguanine was omitted. The second type bound folic acid noncooperatively with Kd = 7 × 10−7 M. Deamino-folic acid did not compete even at a 1000-fold excess. This type may correspond to the chemotactic receptor.  相似文献   

20.
Both enantiomers of 2-benzyl-3-mercaptopropanoic acid were synthesized starting with racemic 3-acetylthio-2-benzylpropanoic acid methyl ester using a kinetic resolution with -chymotrypsin as a key step, and their inhibitory activities against carboxypeptidase A were determined to show that the S-isomer is much more potent (Ki = 7.8 × 10−9 M) than the racemic acid (Ki = 1.1 × 10−8 M).  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