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 734 毫秒
1.
2.
Summary The present study was undertaken to assess and compare the toxic effects of papaverine hydrochloride and its metabolites. Primary cell cultures of rat hepatocytes were treated with papavarine (papaver), 3′-O-desmethyl (3′-OH), 4′-O-desmethyl (4′-OH), and 6-O-desmethyl (6-OH) papaverine at 1×10−5, 1×10−4, and 1×10−3 M for 4,8, 12, and 24-h periods. Cell injury was determined by: a) cell viability using the trypan blue exclusion test; b) cytosolic enzyme leakage of lactate dehydrogenase and aspartate aminotransferase; c) morphologic alterations; and d) lactate: pyruvate (L:P) ratios. Cell cultures showed concentration-and time-dependent responses. For example, a decrease in cell viability and an increase in enzyme leakage were observed after cell treatment with 1×10−4 and 1×10−3 M papaver for 8 h; 1×10−3 M 6-OH papaverine for 8 h and 1×10−4 M for 24 h; and 1×10−3 M 4′-OH papaverine for 24 h (P<0.05). Furthermore, changes in morphology correlated to cell viability and enzyme release in those cultures treated with papaver, 4′-OH and 6-OH papaverine. Some of these changes included size deformation, cell detachment from the dishes, and cell necrosis. On the other hand, an increase in L:P ratios (P<0.05) was detected with papaver as early as 8 h with 1×10−4 and 1×10−3 M and 12 h with 1×10−5 M; 6-OH showed an increase, in L:P ratios at 8 h with 1×10−3 M and 12 h with 1×10−4 M; these changes were evident with 4′-OH at 12 h with 1×10−3 M. In contrast, cells treated with 3′-OH papaverine did not show significant damage with any time period and concentration used in this study. The results of this study indicate that papaverine-derived metabolites are less cytotoxic than its parent compound, papaver. The toxicity was ranked as follows: papaver>6-OH>4′-OH>−3′-OH. This work was supported in part by grant ES04200-02 from the National Institute of Environmental Health Sciences, Bethesda, MD. Presented in part at the fall ASPET meeting in Salt Lake City, August, 1989. Daniel Acosta is a Burroughs Wellcome Scholar in Toxicology.  相似文献   

3.
An extracellular nuclease was purified 165-fold with a specific activity of 41,250 U/mg poly(U) by chromatography with modified chitosan from the culture of marine fungus Penicillium melinii isolated from colonial ascidium collected near Shikotan Island, Sea of Okhotsk, at a depth of 123 m. The purified nuclease is a monomer with the molecular weight of 35 kDa. The enzyme exhibits maximum activity at pH 3.7 for DNA and RNA. The enzyme is stable until 75°C and in the pH range of 2.5–8.0. The enzyme endonucleolytically degrades ssDNA and RNA by 3′–5′ mode to produce 5′-oligonucleotides and 5′-mononucleotides; however, it preferentially degrades poly(U). The enzyme can digest dsDNA in the presence of pregnancy-specific beta-1-glycoprotein-1. The nuclease acts on closed circular double-stranded DNA to produce opened circular DNA and then the linear form DNA by single-strand scission. DNA sequence encoding the marine fungus P. melinii endonuclease revealed homology to S1-type nucleases. The tight correlation found between the extracellular endonuclease activity and the rate of H3-thymidine uptake by actively growing P. melinii cells suggests that this nuclease is required for fulfilling the nucleotide pool of precursors of DNA biosynthesis during the transformation of hyphae into the aerial mycelium and conidia in stressful environmental conditions.  相似文献   

4.
Nuclease Stn α from Streptomyces thermonitrificans hydrolyses DNA and RNA at the rate of approximately 10:l. The optimum pH and temperature for RNA hydrolysis were 7.0 and 45°C. The RNase activity of nuclease Stn α had neither an obligate requirement of metal ions nor was it activated in the presence of metal ions. The enzyme was inhibited by Zn2+, Mg2+, Co2+, and Ca2+; inorganic phosphate; pyrophosphate; NaCl; KCl; and metal chelators. It was stable at high concentrations of urea but susceptible to low concentrations of Sodium dodecyl sulfate and guanidine hydrochloride. The rates by which nuclease Stn α hydrolysed polyribonucleotides occurs in the order of poly A >> RNA >> poly U > poly G > poly C. The enzyme cleaved RNA to 3′ mononucleotides with preferential liberation of 3′AMP, indicating it to be an adenylic acid preferential endonuclease.  相似文献   

5.
From the mycelia of Neurospora crassa (wild type No. 6068) multiple forms of a nuclease which had very close isoelectric points (pI = 9.6 (peak I), 9.4 (peak II)) were isolated by ampholine electrofocusing column chromatography (pH 8.5 ~ 10). The nuclease was about 300-fold purified from the crude extract. The two fractions of Peak I, II were indistinguishable in their enzymatic properties and were considered as manifestation of the same enzyme with minor physicochemical differences. The molecular weight was around 41,000 as estimated by the gel filtration method. The enzyme could hydrolyze both DNA and RNA in the order of heat-denatured DNA > native DNA DNA ≧ RNA. RNA competitively inhibited DNA degradation with this enzyme. The enzyme was therefore regarded as a nuclease. The pH optimum was around pH 6.5 toward native DNA, pH 6.7 toward heat-denatured DNA and pH 7.9 toward RNA. The temperature optimum was around 40°C toward these substrates and most of the activities were lost by heating at 55°C for 15 min. The enzyme required Mg2+ for action toward heat-denatured DNA and Mg2+, Mn2+ or Co2+ toward native DNA. In the presence of EDTA, the activities toward both types of DNA were lost and recovered by addition of the respective activating metallic ions. p-CMB inhibited this nuclease, but β-mercapto-ethanol and glutathione had no effect. Polyamìnes showed no activation of the nuclease for DNA degradation.  相似文献   

6.
The optical activity induced on binding the drug netrospin (NET) in the minor groove of DNA is studied in five oligonucleotides (OGNs) as a function of (1) the size of the binding site in (5′-(GC)2AATT(GC)2-3′)2 (OGN 1a) versus (5′-(GC)2AAATTT(GC)2-3′)2 (OGN 1b) and (2) the distance between two AATT binding sites in (5′-(GC)2AATT(GC) x AATT(GC)2-3′)2, with x = 1, 2, or 3 (OGNs 2a, b, c, respectively). NET binding is monitored via the induced circular dichroism (CD) at ~315 nm, where the nucleic acids are optically inactive. The CD titrations, fit to a tight binding model, yield lower limits for the binding constant, Ka, ≥8 × 107 M−1 for OGN 1a and ≥2 × 108 M−1 for OGNs 2a, b, c in 1 mM buffer. In 100 mM buffer, tight binding occurs in all five OGNs with Ka ≥ 8 × 107 M−1 for OGN 1a and ≥1 × 108 M−1 for OGNs 1b and 2a, b, c. In contrast, the elongated AAATTT binding site of OGN 1b results in weak binding of NET in 1 mM buffer, where competing electrostatic interactions with the solvent environment are lower. In the constructs with two binding sites, the increase in flexibility introduced by intervening GC base pairs does not induce co-operative binding, although differences in the number of binding sites, n (2.05–2.65), indicate that there may be differences in the way NET is bound in OGNs 2a, b, c. In addition, the large shifts in the absorption spectra induced in bound versus free NET, and effects on the CD spectral bands at higher energy, are discussed in terms of electrostatic and excitonic interactions.  相似文献   

7.
The toxicity of iron is believed to originate from the Fenton reaction which produces the hydroxyl radical and/or oxoiron(2+). The effect of nitrogen monoxide on the kinetics of the reaction of iron(II) bound to citrate, ethylenediamine-N,N′-diacetate (edda), ethylenediamine-N,N,N′,N′-tetraacetate (edta), (N-hydroxyethyl)amine-N,N′,N′-triacetate (hedta), and nitrilotriacetate (nta) with hydrogen peroxide was studied by stopped-flow spectrophotometry. Nitrogen monoxide inhibits the Fenton reaction to a large extent. For instance, hydrogen peroxide oxidizes iron(II) citrate with a rate constant of 5.8×103 M−1 s−1, but in the presence of nitrogen monoxide, the rate constant is 2.9×102 M−1 s−1 . Similar to hydrogen peroxide, the reaction of tert-butyl hydroperoxide with iron(II) complexes is also efficiently inhibited by nitrogen monoxide. Generally, nitrogen monoxide binds rapidly to a coordination site of iron(II) occupied by water. The rate of oxidation is influenced by the rate of dissociation of the nitrogen monoxide from iron(II). Electronic Supplementary Material Supplementary material is available for this article at  相似文献   

8.
Relative DNase, RNase (efficiency of hydrolysis of ribo- and deoxyribooligonucleotides (ON)), and phosphatase (removal of the ON 5′ terminal phosphate) catalytic activities of antibodies (AB) obtained after rabbit immunization by DNA, DNase I, and DNase II were compared. It is shown that electrophoretically homogeneous preparations of polyclonal AB from non-immunized rabbits did not exhibit such activities. Immunization of rabbits by DNA, DNase I, and DNase II results in generation of IgG abzymes that exhibit high activity in the ON hydrolysis reaction and even higher activity in cleavage of 5′ terminal phosphate of ON. In this case K m values for supercoiled plasmid DNA and ON found in reactions of their AB-dependent nuclease hydrolysis and phosphatase cleavage of 5′ terminal phosphate differ by 2–4 orders of magnitude. This shows that nuclease and phosphatase activities belong to different abzyme fractions within polyclonal AB. Thus, in this work data indicative of the possibility of a formation of antibodies exhibiting phosphatase activity after immunization of animals with DNA, DNase I, and DNase II, were obtained for the first time. Possible reasons for production of AB with phosphatase activity after immunization of rabbits with these immunogens are discussed.  相似文献   

9.
Summary The ability of Pseudomonas aeruginosa to accumulate Cd(II) ions from wastewater industries was experimentally investigated and mathematically modelled. From the potentiometric titration and non-ideal competitive analysis (NICA) model, it was found that the biomass contains three acidic sites. The values of proton binding (pK i =1.66±3.26×10−3, 1.92±1.63×10−4 and 2.16±3.79×10−4) and binding constant of cadmium metal ions (pK M1=1.99±2.45×10−3 and pK M2=1.67±4.08×10−3) on the whole surface of biomass showed that protonated functional groups and biosorption of Cd(II) ions could be attributed to a monodentate binding to one acidic site, mainly the carboxylic group. From the isothermal sorption experimental data and Langmuir model, it was also found that the value of Langmuir equilibrium (pK f) constant is 2.04±2.1×10−5 suggesting that the carboxyl group is the main active binding site. In addition, results showed that the maximum cadmium capacity (q max) and affinity of biomass towards cadmium metal ions (b) at pH 5.1 and 20 min were 96.5±0.06 mg/g and 3.40×10−3± 2.10×10−3, respectively. Finally, interfering metal ions such as Pb(II), Cu(II), Cr(III), Zn(II), Fe(II), Mn(II), Ca(II) and Mg(II) inhibited Cd(II) uptake. Comparing the biosorption of Cd(II) by various Pseudomonas isolates from contaminated environment samples (soil and sewage treatment plant) showed that maximum capacities and equilibrium times were different, indicating that there was a discrepancy in the chemical composition between biomasses of different strains.  相似文献   

10.
Apyrase and 5′-nucleotidase activities were analyzed in an ethidium bromide (EB) demyelinating model associated with interferon-β (IFN-β). The animals were divided in groups: I, control (saline); II, saline and IFN-β; III, EB and IV, EB and IFN-β. After 7, 15 and 30 days the animals (n=5) were sacrificed and the cerebral cortex was removed for synaptosome preparation and enzymatic assays. Apyrase activity using ATP as substrate increased in groups II, III and IV (P<0.001) after 7 days and in groups III and IV (P<0.001) after 15 days. Using ADP as substrate, an activation of this enzyme was observed in group III (P<0.05) after seven and 15 days. The 5′-nucleotidase activity increased in group III (P<0.05) after 7 days and in groups II, III and IV (P<0.001) after 15 days. After 30 days treatment, no significant alteration was observed in enzyme activities. Results showed that apyrase and 5′-nucleotidase activities are altered in demyelination events and that IFN-β was able to regulate the adenine nucleotide hydrolysis.  相似文献   

11.
 Reaction of [Pt(dien)Cl]+ (1) with the 14-mer oligonucleotide 5′-d(ATACATGGTACATA) (I) gave rise to two major species which corresponded to the 5′-G and 3′-G platinated monofunctional adducts, and a minor amount of the bis-platinated adduct formed during the later stages of the reaction. The reaction of (1) with the related octamer 5′-d(ATACATGG) (II) was also investigated. Kinetic data obtained by HPLC showed that the 5′-G and 3′-G bases of the 14-mer oligonucleotide were platinated at similar rates: the second-order rate constant is 53×10–2 M–1 s–1 at 298 K in 0.1 M NaClO4. However, the platination rate of 5′-G of the octamer (II) (k=69×10–2 M–1 s–1) was enhanced by a factor of three compared to the rate of platination at 3′-G (k=22×10–2 M–1 s–1). All the adducts were separated by HPLC and characterized by NMR spectroscopy, enzymatic digestion and MALDI-TOF mass spectrometry. 1H and 15N NMR shifts suggest that there are distinct conformational differences between 14-mer duplexes platinated at 5′-G (I5′ ds) and 3–G (I3′ ds). Molecular mechanics modelling indicates that rotation around the Pt-N7 bond is more restricted in the case of the 5′-G adduct than in that of the 3′-G adduct. The binding of {Pt(dien)}2+ to 5′-GN7 and 3′-GN7 in the monofunctional adducts of (I) was shown to be reversible upon the addition of high concentrations of chloride ions. Received: 3 July 1998 / Accepted: 10 November 1998  相似文献   

12.
Betaine aldehyde dehydrogenase (BADH; EC 1.2.1.8) is the enzyme that catalyzes the second step in the synthesis of the osmoprotectant, glycine betaine. NAD-dependent BADH was purified from Avena sativa shoots by DEAE Sephacel, hydroxyapatite, 5′-AMP Sepharose 4B, Mono Q and TSK-GEL column chromatographies to homogeneity by the criterion of native PAGE, and the properties of BADH were compared with those of aminoaldehyde dehydrogenase purified to homogeneity from A. sativa. The molecular mass estimated by both gel filtration using TSK-GEL column and Sephacryl S-200 was 120 and 115, kDa, respectively. The enzyme is a homodimer with a subunit molecular mass of 61 kDa as shown by SDS-PAGE. The pI value of the enzyme was found to be 6.3. The purified enzyme catalyzed not only the oxidation of betaine aldehyde (BAL), but also that of aminoaldehydes, 3-aminopropionaldehyde (APAL), 4-aminobutyraldehyde (ABAL), and 4-guanidinobutyraldehyde (GBAL). The K m values for BAL, APAL, ABAL and GBAL were 5×10−6, 5.4×10−7, 2.4×10−5 and 5×10−5 M, respectively. APAL showed substrate inhibition at a concentration of 0.1 mM. A fragment of BADH cleaved by V8 protease shared homology with other plant BADHs. Electronic Publication  相似文献   

13.
A study of sequence homologies in four satellite DNAs of man.   总被引:4,自引:0,他引:4  
Satellites I, II, III and IV (Corneo et al., 1968,1970,1971) have been purified from human male placental DNA. The sequences present in these four DNA components have been characterized by analytical buoyant density, thermal denaturation, DNA reassociation, DNA hybridization and gel electrophoresis coupled with hybridization following either HaeIII or EcoRI restriction endonuclease digestion. Satellites III and IV were found to be virtually indistinguishable by a variety of criteria. Cross-satellite reassociation showed that 40% of the molecules present in satellite III contain sequences that are homologous to 10% of the molecules of either satellite I or satellite II. Reassociated satellite I melts as a single component, as do the hybrid duplexes between satellite I and satellite III. In contrast, reassociated satellites II, III and IV, and the hybrid duplexes formed between satellites II and III and between satellites II and IV, melt as two distinct components with different thermal stabilities.Digestion of satellite III with HaeIII gives rise to a series of fragments whose sizes are 2, 3, 4, 5, 6, 7, 8 and 11 times the size of the smallest 0.17 × 103 basepair fragment, in addition to a 3.4 × 103 base-pair male-specific fragment (Cooke, 1976) and high molecular weight material. The sequences contained in the fragments of the HaeIII ladder are diverged from each other as well as being non-homologous with those of the 3.4 × 103 base-pair and high molecular weight fragments. The latter contain EcoRI recognition sites. Satellite II has a similar pattern of fragments to satellite III following digestion with HaeIII, although it can be distinguished from satellite III on the basis of the products of EcoRI digestion. Satellite I contains neither HaeIII nor EcoRI recognition sites. The cross-satellite homologies of the sequences present in fragments of differing sizes produced by restriction enzyme digestion have also been studied.  相似文献   

14.
The study deals with the changes in metabolic processes of galactose breakdown in tobacco mosaic virus infected tobacco tissues treated with a 1.0 × 10−4 M solution of 2,4-D (2,4-dichlorophenoxyacetic acid). At the time of maximum virus reproduction the plants infected exhibited a considerable increase in galactokinase and galactose-1-phosphate uridyltransferase activities. The activity of both galactose metabolizing enzyme systems investigated was not affected markedly by 2,4-D, but its increase was induced preferentially by virus biosynthesis.  相似文献   

15.
Summary The fungus Aspergillus terreus Thorn var. terreus isolated from an Ecuador soil sample was cultured in liquid and solid media and yielded three main metabolites identified as terreic acid (1), butyrolactone I (2) and lovastatin (3). The natural products as well as three synthetic butyrolactone I derivatives were assessed for antimicrobial activity against Gram-positive and Gram-negative bacteria and fungi as well as for seed germination and seedling growth. Furthermore, the compounds were assessed as inhibitors towards the enzymes acetylcholinesterase, β-glucosidase, and β-glucuronidase. Terreic acid, butyrolactone I, butyrolactone 4′,4′′-diacetate (2.1), and 3′-(3-methylbutyl)-butyrolactone II (2.2) were active towards the phytopathogenic bacteria Erwinia carotovora with IC50 of 5 and 4–18 μg/ml, respectively. Under the same experimental conditions, the IC50 of streptomycin was 1.9 μg/ml. 3′-(3-Methylbutyl)-butyrolactone II was moderately active against Pseudomonas syringae and Botrytis cinerea with IC50 of 21μg/ml and MIC of 15.6 μg/ml, respectively. Butyrolactone I also inhibited germination of the dicot Lactuca sativa with an IC50 of 5 × 10−5 M. The IC50 of reference herbicide acetochlor was 1 × 10−5 M. The effect of 2.2 and 2.3, known as butyrolactone III on Panicum millaceum germination and growth was stronger than that of 2 and 2.1. Reduction of the double bond in the isoprenyl side chain of butyrolactone I increased the antibacterial effect against E. carotovora as well as acetylation. To our best knowledge, this is the first report on the antibacterial effect of butyrolactone derivatives towards Erwinia carotovora and the phytopathogenic fungus Botrytis cinerea. The butyrolactone I derivative 2.2 presented a moderate inhibitory effect against the enzyme acetylcholinesterase with an IC50 of 47 μg/ml. Under the same experimental conditions, the reference inhibitor galanthamine had an IC50 of 3 μg/ml.  相似文献   

16.
Two different heterologous expression systems, microsomal fractions of Saccharomyces cerevisiae and transgenic tobacco plants, were used to investigate the enzymatic activities of flavonoid 3′-hydroxylase (GtF3′H) and flavone synthase II (GtFSII) homologues isolated from gentian petals. Recombinant GtF3′H expressed in yeast showed hydroxylation activities in the 3′ position with several flavonoid substrates, while recombinant GtFSII was able to produce flavone from flavanone. GtF3′ H-expressing transgenic tobacco plants showed a slight increase in anthocyanin content and flower color intensity, and conversion of the flavonol quercetin from kaempferol. On the other hand, GtFSII-expressing plants showed a remarkable reduction in anthocyanin content and flower color intensity, and additional accumulation of flavone, especially luteolin derivatives. We demonstrated that two cytochrome P450s from gentian petals have F3′H and FSII enzymatic activities both in vitro and in vivo, and might therefore be useful in modification of flower color using genetic engineering.  相似文献   

17.
Broken cells ofEscherichia coli contain an enzyme system breaking down cyclic 3′,5′-adenosine monophosphate (Ado-3′,5′-P). The enzyme splitting this nucleotide is located in the supernatant fraction at 20,000 ×g. Some characteristics of the enzyme were studied. In contrast with the animal enzyme theEscherichia coli enzyme is not inhibited by caffeine.  相似文献   

18.
Real-time polymerase chain reaction (PCR) is considered a highly sensitive method for the quantification of microbial organisms in environmental samples. This study was conducted to evaluate real-time PCR with SybrGreen detection as a quantification method for sulfate-reducing bacteria (SRB) in industrial wastewater produced by several chemical industries. We designed four sets of primers and developed standard curves based on genomic DNA of Desulfovibrio vulgaris from pure culture and on plasmids containing dissimilatory sulfate reductase (dsrA) or adenosine-5′-phosphosulfate reductase (apsA) genes of SRB. All the standard curves, two for dsrA and two for apsA genes, had a linear range between 0.95 × 102 and 9.5 × 106 copies/μL and between 1.2 × 103 and 1.2 × 107 copies/μL, respectively. The theoretical copy numbers of the tenfold dilutions of D. vulgaris genomic DNA were best estimated (between 2.7 to 10.5 times higher than theoretical numbers) by the standard curve with DSR1F and RH3-dsr-R primers. To mimic the effect of foreign DNA in environmental samples, serial dilutions of D. vulgaris genomic DNA were mixed with Escherichia coli chromosomal DNA (40 ng per assay). This influenced neither PCR amplification nor the quantification of target DNA. Industrial wastewater was sampled during a 15-month period and analyzed for the presence of SRB, based on dsrA gene amplification. SRB displayed a higher abundance during the summer (about 107–108 targets mL−1) and lower during the winter (about 104–105 targets mL−1). The results indicate that our real-time PCR approach can be used for detection of uncultured SRB and will provide valuable information related to the abundance of SRB in durable environmental samples, such as complex and saline industrial wastewaters.  相似文献   

19.
Kogawa K  Kato N  Kazuma K  Noda N  Suzuki M 《Planta》2007,226(6):1501-1509
A UDP-glucose: anthocyanin 3′,5′-O-glucosyltransferase (UA3′5′GT) (EC 2.4.1.-) was purified from the petals of Clitoria ternatea L. (Phaseoleae), which accumulate polyacylated anthocyanins named ternatins. In the biosynthesis of ternatins, delphinidin 3-O-(6″-O-malonyl)-β-glucoside (1) is first converted to delphinidin 3-O-(6″-O-malonyl)-β-glucoside-3′-O-β-glucoside (2). Then 2 is converted to ternatin C5 (3), which is delphinidin 3-O-(6″-O-malonyl)-β-glucoside-3′,5′-di-O-β-glucoside. UA3′5′GT is responsible for these two steps by transferring two glucosyl groups in a stepwise manner. Its substrate specificity revealed the regioselectivity to the anthocyanin′s 3′- or 5′-OH groups. Its kinetic properties showed comparable k cat values for 1 and 2, suggesting the subequality of these anthocyanins as substrates. However, the apparent K m value for 1 (3.89 × 10−5 M), which is lower than that for 2 (1.38 × 10−4 M), renders the k cat/K m value for 1 smaller, making 1 catalytically more efficient than 2. Although the apparent K m value for UDP-glucose (6.18 × 10−3 M) with saturated 2 is larger than that for UDP-glucose (1.49 × 10−3 M) with saturated 1, the k cat values are almost the same, suggesting the UDP-glucose binding inhibition by 2 as a product. UA3′5′GT turns the product 2 into a substrate possibly by reversing the B-ring of 2 along the C2-C1′ single bond axis so that the 5′-OH group of 2 can point toward the catalytic center. K. Kogawa, N. Kato, K. Kazuma, and N. Noda contributed equally to this work.  相似文献   

20.
TspMI, a thermostable isoschizomer of XmaI from a Thermus sp., has been characterized. The enzyme was purified to homogeneity using Cibacron-Blue 3GA agarose, Heparin agarose, SP sephadex C50, and Mono-Q fast protein liquid chromatography and was found to be a homodimer of 40 kDa. Restriction mapping and run-off sequencing of TspMI-cleaved DNA ends depicted that it cleaved at 5′C/CCGGG3′ to generate a four-base, 5′-CCGG overhang. The enzyme was sensitive to methylation of second and third cytosines in its recognition sequence. TspMI worked optimally at 60°C with 6 mM Mg2+, no Na+/K+, and showed no star activity in the presence of 25% glycerol. The enzyme could efficiently digest the DNA labeled with a higher concentration of YOYO-I (one dye molecule to one nucleotide), making it a useful candidate for real-time imaging experiments. Single molecule interaction between TspMI and λ DNA was studied using total internal reflection fluorescence microscopy. The enzyme survived 30 polymerase chain reaction (PCR) cycles in the presence of 10% glycerol and 0.5 M trehalose without any activity loss and, hence, is suitable for incorporation in restriction-endonuclease-mediated selective-PCR for various applications.Electronic Supplementary Material Supplementary material for this article is available at  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