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.
Escherihica coliumC122::Tn5 cells were γ-radiated (137Cs, 750 Gy, under N2), and lac-constitutive mutants were produced at 36% of the wild-type level (the umC strain was not deficient in spontaneous mutagenesis, and the mutational spectrum determined by sequencing 263 spontaneous lacId mutations was very similar to that for the wild-type strain). The specific nature of the umC strain's partial radiation was determined by sequencing 325 radiation-induced lacId mutations. The yields of radiation-induced mutation classes in the umC strain (as a percentage of the wild-type yield) were: 80% for A · T → G · C transitions, 70% for multi-base additions, 60% for single-base deletions, 53% for A · T → C · G transversions, 36% for G · C → A · T transitions, 25% for multi-base deletions, 21% for A · T → T · A transversions, 11% for G · C → C · G transversions, 9% for G · C → T · A transversions and 0% for multiple mutations. Based on these deficiencies and other factors, it is concluded that the umC strain is near-normal for A · T → G · C transitions, single-base deletions and possibly A · T → C · G transversions; is generally deficient for mutagenesis at G · C sites fro transversions, and is grossly deficient in multiple mutations. Damage at G · C sites seems more difficult for translesion DNA synthesis to bypass than damage at A · T sites, and especially when trying to produced a transversion. The yield of G · C → A · T transitions in the umC strain *36% of the wild-type level) argues that a basic sites are involved in no more than 64% of γ-radiation-induced base substitutions in the wild-type strain. Altogether, these data suggest that the UmuC and UmuD′ proteins facilitate, rather than being absolutely required for, translesion DNA synthesis; with the degree of facilitation being dependent both on the nature of the noncoding DNA damage, i.e., at G · C vs A · T sites, and on the nature of the misincorporated base, i.e., whether it induces transversions or transitions.  相似文献   

2.
Peroxyacetyl nitrate (PAN) is a ubiquitous air pollutant formed from NO(2) reacting with acetoxy radicals generated from ambient aldehydes in the presence of sunlight and ozone. It contributes to eye irritation associated with photochemical smog and is present in most urban air. PAN was generated in a chamber containing open petri dishes of Salmonella TA100 (gas-phase exposure). After subtraction of the background mutation spectrum, the spectrum of PAN-induced mutants selected at 3.1-fold above the background mutant yield was 59% GC-->TA, 29% GC-->AT, 2% GC-->CG, and 10% multiple mutations - primarily GG-->TT tandem-base substitutions. Using computational molecular modeling methods, a mechanism was developed for producing this unusual tandem-base substitution. The mechanism depends on the protonation of PAN near the polyanionic DNA to release NO(2)(+) resulting in intrastrand dimer formation. Insertion of AA opposite the dimerized GG would account for the tandem GG-->TT transversions. Nose-only exposure of Big Blue((R)) mice to PAN at 78ppm (near the MTD) was mutagenic at the lacI gene in the lung (mutant frequency +/-S.E. of 6.16+/-0.58/10(5) for controls versus 8.24+/-0.30/10(5) for PAN, P=0.016). No tandem-base mutations were detected among the 40 lacI mutants sequenced. Dosimetry with 3H-PAN showed that 24h after exposure, 3.9% of the radiolabel was in the nasal tissue, and only 0.3% was in the lung. However, based on the molecular modeling considerations, the labeled portion of the molecule would not have been expected to have been bound covalently to DNA. Our results indicate that PAN is weakly mutagenic in the lungs of mice and in Salmonella and that PAN produces a unique signature mutation (a tandem GG-->TT transversion) in Salmonella that is likely due to a GG intrastrand cross-link. Thus, PAN may pose a mutagenic and possible carcinogenic risk to humans, especially at the high concentrations at which it is present in some urban environments.  相似文献   

3.
We used a double germination phenology or “move-along” experiment (sensu Baskin and Baskin, 2003) to characterize seed dormancy in two medicinal woodland herbs, Collinsonia canadensis L. (Lamiaceae) and Dioscorea villosa L. (Dioscoreaceae). Imbibed seeds of both species were moved through the following two sequences of simulated thermoperiods: (a) 30/15 °C→20/10 °C→15/6 °C→5 °C→15/6 °C→20/10 °C→30/15 °C, and (b) 5 °C→15/6 °C→20/10 °C→30/15 °C→20/10 °C→15/6 °C→5 °C. In each sequence, seeds of both species germinated to high rates (>85%) at cool temperatures (15/6 and 20/10 °C) only if seeds were previously exposed to cold temperatures (5 °C). Seeds kept at four control thermoperiods (5, 15/6, 20/10, 30/15 °C) for 30 d showed little or no germination. Seeds of both species, therefore, have physiological dormancy that is broken by 12 weeks of cold (5 °C) stratification. Morphological studies indicated that embryos of C. canadensis have “investing” embryos at maturity (morphological dormancy absent), whereas embryos of D. villosa are undeveloped at maturity (morphological dormancy present). Because warm temperatures are required for embryo growth and cold stratification breaks physiological dormancy, D. villosa seeds have non-deep simple morphophysiological dormancy (MPD). Neither species afterripened in a 6-month dry storage treatment. Cold stratification treatments of 4 and 8 weeks alleviated dormancy in both species but C. canadensis seeds germinated at slower speeds and lower rates compared to seeds given 12 weeks of cold stratification. In their natural habitat, both species disperse seeds in mid- to late autumn and germinate in the spring after cold winter temperatures alleviate endogenous dormancy.  相似文献   

4.
In order to study the disposition of dimethylamphetamine (DMAP) and its metabolites, DMAP N-oxide, methamphetamine (MA) and amphetamine (AP), from plasma to hair in rats, a simultaneous determination method for these compounds in biological samples using gas chromatography–mass spectrometry with selected ion monitoring (GC–MS-SIM) was developed. As DMAP N-oxide partially degrades to DMAP and MA during GC–MS analysis, it was necessary to avoid conditions which co-extract the N-oxide in the sample preparation so as to assure no contribution of artifactual products from DMAP N-oxide in the detection of the other compounds. For confirmation of the satisfactory separation of DMAP N-oxide from the others, the internal standards used for quantification were labeled with different numbers of deuterium atoms. Determination of unchanged DMAP was performed without any derivatization, that of DMAP N-oxide was carried out after conversion into trifluoroacetyl-MA by reaction with trifluoroacetic anhydride, and MA and AP were quantified after trifluoroacetyl-derivatization.After intraperitoneal administration of DMAP HCl to pigmented hairy rats (5 mg kg−1 day−1, 10 days, n=3), concentrations of DMAP and its metabolites in urine, plasma and hair were measured by GC–MS-SIM. The area under the concentration versus time curves (AUCs) of DMAP, DMAP N-oxide, MA and AP in the plasma were 397.2±97.5, 279.7±68.3, 18.4±1.2 and 15.9±2.2 μg min ml−1, while their concentrations in the hair newly grown for 4 weeks after administration were 4.82±0.67. 0.45±0.09, 3.25±0.36 and 0.89±0.05 ng mg−1, respectively. This fact suggested that the incorporation tendency of DMAP N-oxide from plasma into hair was distinctly low in comparison with the other compounds.  相似文献   

5.
A water-soluble seed gum was isolated from seed endosperm of Cassia javahikai. The acid-catalyzed fragmentation, methylation, selective enzymatic degradation and periodate oxidation suggested a heteropolymeric structure for the polysaccharide. The polysaccharide was shown to have a linear chain of β(1 → 4) linked d-mannopyranosyls units with side chains of α(1 → 6) d-galactopyranosyl units. Grafting of polyacrylamide onto the gum was performed using K2S2O8/ascorbic acid redox system in presence of Ag+ as catalyst at 35 ± 2 °C. The viscosity of the gum solution increased on grafting and the grafted gum was observed to resist biodegradation for more than 256 h. Thermogravimetric analysis revealed that grafted gum was more thermally stable than native gum.  相似文献   

6.
We examined the purine alkaloid content and purine metabolism in cacao (Theobroma cacao L.) plant leaves at various ages: young small leaves (stage I), developing intermediate size leaves (stage II), fully developed leaves (stage III) from flush shoots, and aged leaves (stage IV) from 1-year-old shoots. The major purine alkaloid in stage I leaves was theobromine (4.5 μmol g–1 fresh weight), followed by caffeine (0.75 μmol g–1 fresh weight). More than 75% of purine alkaloids disappeared with subsequent leaf development (stages II–IV). In stage I leaves, 14C-labelled adenine, adenosine, guanine, guanosine, hypoxanthine and inosine were converted to salvage products (nucleotides and nucleic acids), to degradation products (ureides and CO2) and to purine alkaloids (3- and 7-methylxanthine, 7-methylxanthosine and theobromine). In contrast, 14C-labelled xanthine and xanthosine were not used for nucleotide synthesis. They were completely degraded, but nearly 20% of [8-14C]Xanthosine was converted in stage I leaves to purine alkaloids. These observations are consistent with the following biosynthetic pathways for theobromine: (a) AMP → IMP → 5′-xanthosine monophosphate → xanthosine → 7-methylxanthosine → 7-methylxanthine → theobromine; (b) GMP → guanosine → xanthosine → 7-methylxanthosine → 7-methylxanthine → theobromine; (c) xanthine → 3-methylxanthine → theobromine. Although no caffeine biosynthesis from 14C-labelled purine bases and nucleosides was observed during 18 h incubations, exogenously supplied [8-14C]Theobromine was converted to caffeine in young leaves. Conversion of theobromine to caffeine may, therefore, be slow in cacao leaves. No purine alkaloid synthesis was observed in the subsequent growth stages (stages II–IV). Significant degradation of purine alkaloids was found in leaves of stages II and III, in which [8-14C]Theobromine was degraded to CO2 via 3-methylxanthine, xanthine and allantoic acid. [8-14C]Caffeine was catabolised to CO2 via theophylline (1,3-dimethylxanthine) or theobromine.  相似文献   

7.
Previously, we have demonstrated that cadmium acetate significantly induces hprt mutation frequency in Chinese hamster ovary (CHO)-K1 and that 3-amino-1,2,4-triazole (3AT), a catalase inhibitor, potentiates the mutagenicity of cadmium [Chem. Res. Toxicol. 9 (1996) 1360–1367]. In this study, we investigate the role of intracellular peroxide in the molecular nature of mutations induced by cadmium. Using 2′,7′-dichlorofluorescin diacetate and fluorescence spectrophotometry, we have shown that cadmium dose-dependently increased the amounts of intracellular peroxide and the levels were significantly enhanced by 3AT. Furthermore, we have characterized and compared the hprt mutation spectra in 6-thioguanine-resistant mutants derived from CHO-K1 cells exposed to 4 μM of cadmium acetate for 4 h in the absence and presence of 3AT. The mutation frequency induced by cadmium and cadmium plus 3AT was 11- and 16-fold higher than that observed in untreated populations (2.2×10−6), respectively. A total of 40 and 51 independent hprt mutants were isolated from cadmium and cadmium plus 3AT treatments for mRNA-polymerase chain reaction (PCR), genomic DNA-PCR and DNA sequencing analyses. 3AT co-administration significantly enhanced the frequency of deletions induced by cadmium. Cadmium induced more transversions than transitions. In contrast, 3AT co-administration increased the frequency of GC→AT transitions and decreased the frequencies of TA→AT and TA→GC transversions. Together, the results suggest that intracellular catalase is important to prevent the formation of oxidative DNA damage as well as deletions and GC→AT transitions upon cadmium exposure.  相似文献   

8.
We report here the development and validation of an LC–MS method for quantitation of loperamide (LOP) and its N-demethyl metabolite (DMLOP) in human plasma. O-Acetyl-loperamide (A-LOP) was synthesized by us for use as an internal standard in the assay. After addition of the internal standard, the compounds of interest were extracted with methyl tert.-butylether and separated by HPLC on a C18 reversed-phase column using an acetonitrile–water gradient containing 20 mM ammonium acetate. The three compounds were well separated by HPLC and no interfering peaks were detected at the usual concentrations found in plasma. Analytes were quantitated using positive electrospray ionization in a triple quadrupole mass spectrometer operating in the MS–MS mode. Selected reaction monitoring was used to quantify LOP (m/z 477→266), DMLOP (m/z 463→252) and A-LOP (m/z 519→266) on ions formed by loss of the 4-(p-chlorophenyl)-4-hydroxy-piperidyl group upon low energy collision-induced dissociation. Calibration curves, which were linear over the range 1.04 to 41.7 pmol/ml (LOP) and 1.55 to 41.9 pmol/ml (DMLOP), were run contemporaneously with each batch of samples, along with low (4.2 pmol/ml), medium (16.7 pmol/ml) and high (33.4 pmol/ml) quality control samples. The lower limit of quantitation (LLQ) of LOP and DMLOP was about 0.25 pmol/ml in plasma. The extraction efficiency of LOP and DMLOP from human plasma was 72.3±1.50% (range: 70.7–73.7%) and 79.4±12.8% (64.9–88.8%), respectively. The intra- and inter-assay variability of LOP and DMLOP ranged from 2.1 to 14.5% for the low, medium and high quality control samples. The method has been used successfully to study loperamide pharmacokinetics in adult humans.  相似文献   

9.
A sensitive, selective, and reproducible GC–MS–SIM method was developed for determination of artemether (ARM) and dihydroartemisinin (DHA) in plasma using artemisinin (ART) as internal standard. Solid phase extraction was performed using C18 Bond Elut cartridges. The analysis was carried out using a HP-5MS 5% phenylmethylsiloxane capillary column. The recoveries of ARM, DHA and ART were 94.9±1.6%, 92.2±4.1% and 81.3±1.2%, respectively. The limit of quantification in plasma was 5 ng/ml (C.V.≤17.4% for ARM and 15.2% for DHA). Calibration curves were linear with R2≥0.988. Within day coefficients of variation were 3–10.4% for ARM and 7.7–14.5% for DHA. Between day coefficients of variations were 6.5–15.4% and 7.6–14.1% for ARM and DHA. The method is currently being used for pharmacokinetic studies. Preliminary data on pharmacokinetics showed Cmax of 245.2 and 35.6 ng/ml reached at 2 and 3 h and AUC0–8h of 2463.6 and 111.8 ngh/ml for ARM and DHA, respectively.  相似文献   

10.
Live, intact third-stage larvae (L3s) of Strongyloides ratti in the absence of exogenous substrates consumed oxygen at a rate (E-QO2) of 181.8 ± 12.4 ng atoms min−1 mg dry weight−1 at 35°C. Respiratory electron transport (RET) Complex I inhibitor rotenone (2 μm) produced 33 ± 6.5% inhibition of the E-QO2. Unusually the rotenone-induced inhibition was not relieved by 5 μm-succinate. The E-QO2 of intact L3s was refractory to RET Complex III inhibitor antimycin A at 2 μm; 4 μm-antimycin inhibited ≤ 10% of the E-QO2. The electron donor couple ascorbate/TMPD augmented the E-QO2 in the presence of rotenone (2 μm) and antimycin A (4 μm) by 110%. Azide (1 mm) stimulated the antimycin A refractory QO2 by 36.6 ± 7.2% which was only partially inhibited by 1.0 mm-KCN (IC50 = 0.8 mm). The data suggest the presence of classical (CPW) and alternate (APW) electron transport pathways in S. ratti L3s.  相似文献   

11.
The kinetic properties of a microsomal gill (Na+,K+)-ATPase from the blue crab Callinectes danae were analyzed using the substrate p-nitrophenylphosphate. The (Na+,K+)-ATPase hydrolyzed PNPP obeying cooperative kinetics (n=1.5) at a rate of V=125.4±7.5 U mg−1 with K0.5=1.2±0.1 mmol l−1; stimulation by potassium (V=121.0±6.1 U mg−1; K0.5=2.1±0.1 mmol l−1) and magnesium ions (V=125.3±6.3 U mg−1; K0.5=1.0±0.1 mmol l−1) was cooperative. Ammonium ions also stimulated the enzyme through site–site interactions (nH=2.7) to a rate of V=126.1±4.8 U mg−1 with K0.5=13.7±0.5 mmol l−1. However, K+-phosphatase activity was not stimulated further by K+ plus NH4+ ions. Sodium ions (KI=36.7±1.7 mmol l−1), ouabain (KI=830.3±42.5 μmol l−1) and orthovanadate (KI=34.0±1.4 nmol l−1) completely inhibited K+-phosphatase activity. The competitive inhibition by ATP (KI=57.2±2.6 μmol l−1) of PNPPase activity suggests that both substrates are hydrolyzed at the same site on the enzyme. These data reveal that the K+-phosphatase activity corresponds strictly to a (Na+,K+)-ATPase in C. danae gill tissue. This is the first known kinetic characterization of K+-phosphatase activity in the portunid crab C. danae and should provide a useful tool for comparative studies.  相似文献   

12.
A GC method using a novel derivatization reagent, 2′,2′,2-trifluoroethyl chloroformate (TFECF), for the derivatization of primary and secondary aliphatic amines with the formation of carbamate esters is presented. The method is based on a derivatization procedure in a two-phase system, where the carbamate ester is formed. The method is applied to the determination of 1,6-hexamethylene diamine (HDA) in aqueous solutions and human urine, using capillary GC. Detection was performed using thermionic specific detection (TSD) and mass spectrometry (MS)—selective-ion monitoring (SIM) using electron-impact (EI) and chemical ionization (CI) with ammonia monitoring both positive (CI)+ and negative ions (CI). Quantitative measurements were made in the chemical ionization mode monitoring both positive and negative ions. Tetra-deuterium-labelled HDA (TDHDA; H2NC2H2(CH2)4C2H2NH2) was used as the internal standard for the GC—MS analysis. In CI+ the m/z 386 and the m/z 390 ions corresponding to the [M + 18]+ ions (M = molecular ion) of HDA—TFECF and TDHDA—TFECF were measured; in CI the m/z 267 and the m/z 271 ions corresponding to the [M — 101] ions. The overall recovery was found to be 97 ± 5% for a HDA concentration of 1000 μg/l in urine. The minimal detectable concentration in urine was found to be less than 20 μg/l using GC—TSD and 0.5 μg/l using GC—SIM. The overall precision for the work-up procedure and GC analysis was ca. 3% (n = 5) for 1000 μg/l HDA-spiked urine, and ca. 4% (n = 5) for 100 μg/l. The precision using GC—SIM for urine samples spiked to a concentration of 5 μg/l was found to be 6.3% (n = 10).  相似文献   

13.
The heart rate response to isoproterenol (HR-Iso), density and affinity (kd) of β-adrenergic (β-AR) and muscarinic (M2) receptors were compared among three rodents with different generation-life histories of confinement and of high altitude exposure. The European guinea pig (Cavia porcellus) (EGp), a laboratory animal that arrived in Europe after the Spanish Conquest of South America and the Peruvian guinea pig (C. porcellus) (PGp), a semi-wild animal that came from the altiplano to sea level at least 25 generations ago, were used for intra-species comparison. Wistar rats (WR) were used for inter-species comparison as representative of a typical sea level laboratory animal. The HR-Iso was lower in EGp than in the PGp. The PGp showed the highest β-AR density (P<0.0005) and the highest β-AR kd values (P<0.0005) when compared to both EGp and WR groups (β-AR Bmax (fmol mg−1 prot), WR, 19±4; Egp, 34±10; PGp, 74±15. β-AR kd (pM), WR, 24±10; Egp, 17±7; PGp, 39±14). In contrast, PGp showed lower M2 receptor density values than the EGp (P<0.0005). The WR had the highest M2 receptor densities (M2 Bmax (fmol mg−1 prot), WR, 188±15; Egp, 147±9; PGp, 118±6 and M2 kd (pM), WR, 65±12; Egp, 67±6; PGp, 92±2). The inter and intra-species differences found may be related to their respective history of confinement rather than to their history of exposure to high altitude.  相似文献   

14.
The simultaneous transplantation and intrathymic tolerance induction (STITTI) protocol induces a longlasting state of functional tolerance in over 90% of AO (RT1u) recipients transplanted with a fully MHC-incompatible PVG (RT1c) cardiac allograft. Similar results are obtained when using LEWIS (RT11) rats as recipients of either PVG or DA (RT1avl) grafts. However, when STITTI is performed on PVG and BN (RT1n) as recipient animals receiving spleen cells intrathymically and a cardiac allograft from respectively AO and PVG rats, this procedure results in significantly shorter graft survival (MST PVG → BN 25 ± 9 days; AO → PVG 31 ± 8 days) as compared to the combinations using AO (MST PVG → AO > 236 ± 28 days) and LEWIS (MST PVG → LEW > 366 ± 51 days; DA → LEW > 123 ± 33 days) rats as recipients. Since both PVG and BN rats are relatively deficient in their ability to produce IFNγ and intrathymic IFNγ responses are very dominant upon intrathymic injection of alloantigens, it is argued that the inability to effectively induce a longlasting state of functional tolerance in BN and PVG rats using the STITTI protocol may be related to their decreased IFNγ-production potential.  相似文献   

15.
We studied the seasonal variation on aerobic metabolism and the response of oxidative stress parameters in the digestive glands of the subpolar limpet Nacella (P.) magellanica. Sampling was carried out from July (winter) 2002 to July 2003 in Beagle Channel, Tierra del Fuego, Argentina. Whole animal respiration rates increased in early spring as the animals spawned and remained elevated throughout summer and fall (winter: 0.09 ± 0.02 μmol O2 h− 1 g− 1; summer: 0.31 ± 0.06 μmol O2 h− 1 g− 1). Oxidative stress was assessed at the hydrophilic level as the ascorbyl radical content / ascorbate content ratio (A / AH). The A / AH ratio showed minimum values in winter (3.7 ± 0.2 10− 5 AU) and increased in summer (18 ± 5 10− 5 AU). A similar pattern was observed for lipid radical content (122 ± 29 pmol mg− 1 fresh mass [FW] in winter and 314 ± 45 pmol mg− 1 FW in summer), iron content (0.99 ± 0.07 and 2.7 ± 0.6 nmol mg− 1 FW in winter and summer, respectively) and catalase activity (2.9 ± 0.2 and 7 ± 1 U mg− 1 FW in winter and summer, respectively). Since nitrogen derived radicals are thought to be critically involved in oxidative metabolism in cells, nitric oxide content was measured and a significant difference in the content of the Fe–MGD–NO adduct in digestive glands from winter and summer animals was observed. Together, the data indicate that both oxygen and nitrogen radical generation rates in N. (P.) magellanica are strongly dependent on season.  相似文献   

16.
PAH (N-(4-aminobenzoyl)glycin) clearance measurements have been used for 50 years in clinical research for the determination of renal plasma flow. The quantitation of PAH in plasma or urine is generally performed by colorimetric method after diazotation reaction but the measurements must be corrected for the unspecific residual response observed in blank plasma. We have developed a HPLC method to specifically determine PAH and its metabolite NAc-PAH using a gradient elution ion-pair reversed-phase chromatography with UV detection at 273 and 265 nm, respectively. The separations were performed at room temperature on a ChromCart® (125 mm×4 mm I.D.) Nucleosil 100-5 μm C18 AB cartridge column, using a gradient elution of MeOH–buffer pH 3.9 1:99→15:85 over 15 min. The pH 3.9 buffered aqueous solution consisted in a mixture of 375 ml sodium citrate–citric acid solution (21.01 g citric acid and 8.0 g NaOH per liter), added up with 2.7 ml H3PO4 85%, 1.0 g of sodium heptanesulfonate and completed ad 1000 ml with ultrapure water. The N-acetyltransferase activity does not seem to notably affect PAH clearances, although NAc-PAH represents 10.2±2.7% of PAH excreted unchanged in 12 healthy subjects. The performance of the HPLC and the colorimetric method have been compared using urine and plasma samples collected from healthy volunteers. Good correlations (r=0.94 and 0.97, for plasma and urine, respectively) are found between the results obtained with both techniques. However, the colorimetric method gives higher concentrations of PAH in urine and lower concentrations in plasma than those determined by HPLC. Hence, both renal (ClR) and systemic (ClS) clearances are systematically higher (35.1 and 17.8%, respectively) with the colorimetric method. The fraction of PAH excreted by the kidney ClR/ClS calculated from HPLC data (n=143) is, as expected, always <1 (mean=0.73±0.11), whereas the colorimetric method gives a mean extraction ratio of 0.87±0.13 implying some unphysiological values (>1). In conclusion, HPLC not only enables the simultaneous quantitation of PAH and NAc-PAH, but may also provide more accurate and precise PAH clearance measurements.  相似文献   

17.
A novel polysaccharide designated EPS-1A with an average molecular weight around 40 kDa was fractionated and purified by anion-exchange and gel-filtration chromatography from the crude exopolysaccharide (EPS) isolated from fermentation broth of Cs-HK1, a Tolypocladium sp. fungus isolated from wild Cordyceps sinensis. The structural characteristics of EPS-1A were determined with various methods (e.g. GC, GC–MS, FT-IR, 1H NMR and 13C NMR) and through acid hydrolysis, methylation, periodate-oxidation and Smith degradation. The results suggested that EPS-1A was composed of glucose, mannose and galactose at 15.2:3.6:1.0 M ratio. EPS-1A was a slightly branched polysaccharide and its backbone was composed of (1 → 6)-α-d-glucose residues (77%) and (1 → 6)-α-d-mannose residues (23%). Branching occurred at O-3 position of (1 → 6)-α-d-mannose residues of the backbone with (1 → 6)-α-d-mannose residues and (1 → 6)-α-d-glucose residues, and terminated with β-d-galactose residues.  相似文献   

18.
The fragile histidine triad (Fhit) protein is a homodimeric protein with diadenosine 5′,5-P1,P3-triphosphate (Ap3A) asymmetrical hydrolase activity. We have cloned the human cDNA Fhit in the pPROEX-1 vector and expressed with high yield in Escherichia coli with the sequence Met-Gly-His6-Asp-Tyr-Asp-Ile-Pro-Thr-Thr followed by a rTEV protease cleavage site, denoted as “H6TV,” fused to the N-terminus of Fhit. Expression of H6TV–Fhit in BL21(DE3) cells for 3 h at 37°C produced 30 mg of H6TV–Fhit from 1 L of cell culture (4 g of cells). The H6TV–Fhit protein was purified to homogeneity in a single step, with a yield of 80%, using nickel-nitrilotriacetate resin and imidazole buffer as eluting agent. Incubation of H6TV–Fhit with rTEV protease at 4°C for 24 h resulted in complete cleavage of the H6TV peptide. There were no unspecific cleavage products. The purified Fhit protein could be stored for 3 weeks at 4°C without loss of activity. The pure protein was stable at −20°C for at least 18 months when stored in buffer containing 25% glycerol. Purified Fhit was highly active, with a Km value for Ap3A of 0.9 μM and a kcat(monomer) value of 7.2 ± 1.6 s−1 (n = 5). The catalytic properties of unconjugated Fhit protein and the H6TV–Fhit fusion protein were essentially identical. This indicates that the 24-amino-acid peptide containing the six histidines fused to the N-terminus of Fhit does not interfere in forming the active homodimers or in the binding of Ap3A.  相似文献   

19.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

20.
We measured the toxicity and mutagenicity induced in human diploid lymphoblasts by various radiation doses of X-rays and two internal emitters. [125I]iododeoxyuridine ([125I]dUrd) and [3H]thymidine ([3H]TdR), incorporated into cellular DNA. [125I]dUrd was more effective than [3H]TdR at killing cells and producing mutations to 6-thioguanine resistance (6TGR). No ouabain-resistant mutants were induced by any of these agents. Expressing dose as total disintegrations per cell (dpc), the D0 for cell killing for [125I]dUrd was 28 dpc and for [3H]TdR was 385 dpc. The D0 for X-rays was 48 rad at 37°C. The slopes of the mutation curves were approximately 75 × 10−8 6TGR mutants per cell per disintegration for [125I]dUrd and 2 × 10−8 for [3H]TdR. X-Rays induced 8 × 10−8 6TGR mutants per cell per rad. Normalizing for survival, [125I]dUrd remained much more mutagenic at low doses (high survival levels) than the other two agents. Treatment of the cells at either 37°C or while frozen at −70°C yielded no difference in cytotoxicity or mutation for [125I]dUrd or [3H]TdR, whereas X-rays were 6 times less effective in killing cells at −70°C.Assuming that incorporation was random throughout the genome, the mutagenic efficiencies of the radionuclides could be calculated by dividing the mutation rate by the level of incorporation. If the effective target size of the 6TGR locus is 1000–3000 base pairs, then the mutagenic efficiency of [125I]dUrd is 1.0–3.0 and of [3H]TdR is 0.02–0.06 total genomic mutations per cell per disintegration. 125I disintegrations are known to produce localized DNA double-strand breaks. If these breaks are potentially lethal lesions, they must be repaired, since the mean lethal dose (D0) was 28 dpc. The observations that a single dpc has a high probability of producing a mutation (mutagenic efficiency 1.0–3.0) would suggest, however, that this repair is extremely error-prone. If the breaks need not be repaired to permit survival, then lethal lesions are a subset of or are completely different from mutagenic lesions.  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