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.
We have investigated the effect of chloroform on the phosphorylation of myelin basic proteins because tumor-promoting agents such as phorbol esters and chloroform are known to enhance the activity of protein kinase C. We report that the presence of chloroform, at a concentration known to enhance protein kinase C activity, stimulated the phosphorylation of myelin basic proteins 15-17 fold over control conditions. The phosphorylation of a 50 kiloDalton myelin protein was also stimulated but to a lesser extent. The concentration of chloroform required for the maximal phosphorylation of myelin basic proteins and the 50 kiloDalton protein was approximately 2% (v/v).  相似文献   

2.
Characterization of poly-3-hydroxybutyric acid (PHB) and poly-3-hydroxybutyric-co-valeric acid (PHBV, 13% valerate) in chloroform was performed using size exclusion chromatography coupled to a multi-angle light scattering detector (SEC-MALS). Absolute molar mass averages, molar mass distribution, and the radius of gyration were determined. Three sample preparation methods were examined: dissolution in chloroform (1) at room temperature, (2) at 60 degrees C, and (3) after thermal pretreatment of samples (annealing at 180 degrees C with subsequent quenching in liquid nitrogen). Dissolution at 60 degrees C and dissolution of thermally pretreated samples gave molecularly dissolved PHB and PHBV. At 60 degrees C using acid free chloroform, there was no indication of degradation for up to 120 min dissolution time, whereas thermal degradation of polymers did take place during annealing at 180 degrees C. The degradation rate constants for number and weight average degree of polymerization at 180 degrees C were slightly higher for PHB (5.19 x 10(-5) min(-1), 4.95 x 10(-5) min(-1)) than for PHBV (4.99 x 10(-5) min(-1), 4.54 x 10(-5) min(-1)). The dependence of the radii of gyration on molar mass showed that both polymers form random coils in chloroform. The relationship between the absolute molar masses and relative SEC results was determined. DSC and NMR characterization also gave evidence of the progress of degradation.  相似文献   

3.
The major interaction site for tumor-promoting phorbol esters is the calcium-activated, phospholipid-dependent protein kinase (protein kinase C), a key-element in signal transduction. Binding of phorbol esters results in enzyme activation which mediates, at least in part, the action of these agents. We have investigated the effects of tumor promoter chloroform on protein kinase C activity. Like thrombin and 12-O-tetradecanoylphorbol-13-acetate (TPA), chloroform was able to activate protein kinase C in intact rabbit platelets. In addition, chloroform stimulated enzyme activity as well as TPA binding capacity in cell-free system. Scatchard analysis of the data has shown that chloroform increased the number of phorbol ester binding sites. Structurally related compounds, carbon tetrachloride and methylene chloride, activated the enzyme similarly.  相似文献   

4.
The solubility of phosphatidylcholine (PC) was studied by the spectroscopic analysis and the measurement of the solubility. The qualitative analysis of infrared absorptionspectra confirmed the existence of two types of hydrogen bondings between chloroform and PC, one between chloroform and the CO group of PC and the other between chloroform and the phosphorylcholine group of PC. The quantitative analysiss of the C6 stretching vibration bands of the chloroform-d solution of PC showed that the latter hydrogen bonding mainly contributes to the solubility and that PC dissolves in chloroform to form a complex consisting of a few or more molecules of chloroform and one molecule of PC. We discussed in this report about the molecular organization of PC in chloroform solution.  相似文献   

5.
The covalent binding of chemical carcinogens and mutagens to hemoglobin has been proposed as a dose monitor for environmental exposure. The binding of chloroform to hemoglobin in rats was demonstrated to result from the formation of addition adducts to amino acids in the globin. The altered amino acids were isolated with an amino acid analyzer employing ion exchange chromatography. The covalent binding of orally-administered [14C]chloroform to rat hemoglobin reached a peak within 10 h. Afterwards, the amount of chloroform bound decreased slowly with half remaining at 7 weeks. The extent of chloroform binding was linearly dependent upon dose between 0.1 and 100 μmol/kg. Above 100 μmol/kg chloroform the binding to hemoglobin increased at a reduced rate. The extent of [14C]chloroform binding resulting from either 10 daily doses of 0.1 μmol/kg or a single 1.0 μmol/kg was similar. The binding to hemoglobin in three strains of mice and rats varied from 85 ± 7 μmol/g Hb for Swiss (CFN) mice to 152 ± 14 for Fisher-344 rats. We have demonstrated that the binding of chloroform to hemoglobin appears to have the following attributes of a dose monitor: (1) easily accessible from laboratory animals (and humans); (2) dose dependent; (3) stability, so that exposure can be determined weeks after it has ceased; (4) integration of low exposures which individually would be undetectable.  相似文献   

6.
The formation of14CO2 from 3 μg l−1 labelled chloroform was studied in anaerobic Dutch river sediments. All incubations were performed under anaerobic conditions. The observed first order mineralization kinetics showed half-lives of 2–37 days at 20°C in 12 muddy sediments. In contrast most of the sandy sediment samples did not show a mineralization of chloroform. Most probable number analysis revealed about 3.104 chloroform mineralizing bacteria per g of dry sediment in a muddy sediment and 1–2.103 chloroform mineralizing bacteria per g of dry sediment in a sandy sediment. Therefore the persistence of chloroform in sandy sediments is not caused by the absence of chloroform mineralizing bacteria but by the inactivity of these bacteria. This inactivity of the sandy sediments might allow chloroform from infiltrating river water to reach the groundwater. Mud samples from a relatively unpolluted site showed a similar chloroform mineralization rate compared with the polluted sediments from the rivers Rhine and Meuse. The data indicate that the reductive dechlorination of aliphatic compounds is not influenced at the polluted sites.  相似文献   

7.
Summary Recovery of poly-3-hydroxybutyrate (PHB) in three chlorinated solvents with or without acetone pretreatment and degradation of extracted PHB (99% pure) in hot chloroform were studied. When lyophilized Alcaligenes eutrophus biomass was used, the best results were obtained with acetone pretreatment and solvent reflux for 15 min in methylene chloride or chloroform. Recovered PHB had a 95% purity and molecular weights (Mw) of 1,050,000 and 930,000 g/mol respectively. Further heating resulted in a serious Mw, loss at reflux temperatures. Degradation of extracted PHB at 110°C in chloroform was due to random and chain-end scission, the former being predominant.  相似文献   

8.
Cholesterol oleate with the 13C-label in oleic acid at the carbonyl and/or in the sterol ring at position 4 was synthesized by two methods: (1) cholesterol was condensed with oleic anhydride, prepared from [1-13C] oleic acid, in the presence of dimethylaminopyridine (DMAP) in anhydrous chloroform at room temperature for 4--5 h; (2) cholesterol or 13C-enriched cholesterol at position 4 were reacted with 90% [1-13C]-oleic acid in the presence of dicyclohexylcarbodiimide (DCC) and DMAP at room temperature in anhydrous chloroform for 1.25 h. The single-13C and double-13C-labeled cholesterol oleate were obtained in 90% yields after purification by silicic acid column chromatography. Their purity was assessed by thin-layer chromatography (TLC), high-performance liquid chromatography (HPLC) and 13C-NMR spectroscopy. Tritium-labeled cholesterol oleate was also synthesized by method 1 using the fatty acid anhydride.  相似文献   

9.
The enthalpy of hydrogen-bond formation between guanine (G) and cytosine (C) in o-dichlorobenzene and in chloroform at 25°C has been determined by direct calorimetric measurement. We derivatized 2′-deoxyguanosine and 2′-deoxycytidine at the 5′- and 3′-hydroxyls with triisopropylsilyl groups; these groups increase the solubility of the nucleic acid bases in nonaqueous solvents. Such derivatization also prevents the ribose hydroxyls from forming hydrogen bonds. Consequently, hydrogen-bond formation in our system is primarily between the bases, and to a lesser extent, between base and solvent, and can be measured directly with calorimetry. To obtain the data on base-pair formation, we first took into account the contributions from self-association of each base, and where possible, have determined the ΔH of self-association. From isoperibolic titration calorimetry, our measured ΔH of C2 formation in chloroform is ?1.7 kcal/mol of C. Our measured ΔH of C:G base-pair formation in o-dichlorobenzene is ?6.65 ± 0.32 kcal/mol. Since o-dichlorobenzene does not form hydrogen bonds, the ΔH of C:G base-pair formation in this solvent represents the ΔH of the hydrogen-bonding interaction of C with G in a nonassociating solvent. In contrast, our measured ΔH of C:G base-pair formation in chloroform is ?5.77 ± 0.20 kcal/mol; thus, the absolute value of the enthalpy of hydrogen bonding in the C:G base pair is greater in o-dichlorobenzene than in chloroform. Since chloroform is a solvent known to form hydrogen bonds, the decrease in enthalpic contribution to C:G base pairing in chloroform is due to the formation of hydrogen bonds between the bases and the solvent. The ΔH of hydrogen bonding of G with C reported here differs from previous indirect estimates: Our measurements indicate the ΔH is 50% less in magnitude than the ΔH based on spectroscopic measurements of the extent of interaction. We have also observed that the enthalpy of hydrogen bonding of C with G in chloroform is greater when G is in excess than when C is in excess. This increased heat is due to the formation of C:Gn > 1 complexes that we have observed using 1H-nmr. Although C:G2 structures have previously been observed in triple-stranded polymeric nucleic acids, higher order structures have not been observed between C and G monomers in nonaqueous solvents until now. By using monomers as a model system to investigate hydrogen-bonding interactions in DNA and RNA, we have obtained the following results: A direct measurement of the ΔH of hydrogen bonding in the C:G complex in two nonaqueous solvents, and the first observation of C:Gn > 1 complexes between monomers. These results reinforce the importance of hydrogen bonding in the stabilization of various nucleic acid secondary and tertiary structures.  相似文献   

10.
Fourteen medicinal plants belonging to 13 families were collected and extracted with petroleum ether (PE), chloroform, methanol and water to yield 60 crude extracts. Using agar diffusion method, these extracts were evaluated for antifungal activity on the growth of five phytopathogenic fungi. Among all the extracts tested, PE, chloroform and methanol extracts of Piper betle L. and PE and chloroform extracts of Allamanda cathartica exhibited promising antifungal activity. Minimum inhibitory concentration (MIC) values of the above promising extracts were determined using broth dilution technique and observed that chloroform extract of P. betle L. exhibited the least MIC value ranging from 280 to 1130 μg ml?1. In this study, we report chloroform extract of P. betle L. to be thermally stable even when steam sterilised for the first time and that it could be stored at 4°C with almost no change in its activity for a period of 180 days.  相似文献   

11.
Bioefficacy of hexane, chloroform and ethyl acetate extracts of the seeds of Ceasalpinea bonduc (L.) Roxb. was studied against third instar larvae of Spodoptera litura at 0.25, 0.5, 1, 2.5 and 5.0% concentrations. Significant antifeedant, larvicidal and pupicidal activities and least LC50 values were observed in chloroform extract. The chloroform extract was subjected to fractionation using silica gel column chromatography. Six fractions were obtained; among these, the third fraction showed high antifeedant, larvicidal and pupicidal activities at 1000 ppm concentration. Abnormalities in adults were also observed. All the activities were concentration dependent. C. bonduc could be useful in integrated pest management programme.  相似文献   

12.
Interaction of fd or M13 filamentous phage with a chloroform/water interface induces morphological change, contracting the filaments sequentially into shortened rods (I-forms), and then into spheroidal particles (S-forms). To further investigate this phage contraction, 34 and 26 chloroform-resistant isolates of fd and M13, respectively, were selected after chloroform treatment of wild-type phages at pH 8. 2 and 4 degrees C. DNA sequencing of gene VIII of the 34 fd isolates revealed five different mutants: these were D5H, M28L, V31L, I37T, and S50T. All 26 M13 isolates were I37T. These mutants exhibited variable sensitivity to chloroform, but all contracted much more slowly than wild-type phage during treatment at 4 degrees C. They all contracted like wild-type phage at 37 degrees C. Site-directed mutagenesis showed that the indicated single mutations carried the chloroform resistance. In structural models of the phage, the D5H locus is on the outside and the S50T locus is on the inside. The M28L and I37T loci are buried in a mostly hydrophobic region in the middle. Although these four mutants are spread out radially, they are localized in the axial direction into a thin disk in the model. The last mutant locus, V31L, is out of this disk, but this locus is proximal to the M28L and I37T loci and also in contact with the surface via a deep hydrophobic hole or depression. These five mutants, their locations, and their variable affects on contraction suggest that chloroform-induced contraction involves a specific mechanism rather than a generalized solvent-induced denaturation and that the critical structural changes occur in a localized level in the phage. These results add weight to suggestions that the sequential contraction of filaments-->I-forms-->S-forms mimic corresponding steps in phage penetration, and, in the reverse order, for phage assembly.  相似文献   

13.
The effect of microbial composition on the methanogenic degradation of cellulose was studied using two lines of anaerobic cellulose-fermenting methanogenic microbial cultures at two different temperatures: that at 15 degrees C being dominated by Methanosaeta and that at 30 degrees C by Methanosarcina. In both cultures, CH4 production and acetate consumption were completely inhibited by either 2-bromoethanesulfonate or chloroform, whereas H2 consumption was only inhibited by chloroform, suggesting that homoacetogens utilized H2 concomitantly with methanogens. Hydrogen was the intermediate that was consumed first, while acetate continued to accumulate. At 15 degrees C, acetoclastic methanogenesis smoothly followed H2-dependent CH4 production. Fluorescence in situ hybridization showed that populations of Methanosaeta steadily increased with time from 5 to 25% of total cell counts. At 30 degrees C, two phases of CH4 production were obtained, with acetate consumed after the abrupt increase of Methanosarcina from 0 to 45% of total cell counts. Whereas populations of Methanosaeta were able to adapt after transfer from 15 to 30 degrees C, those of Methanosarcina were not, irrespective of during which phase the cultures were transferred from 30 degrees C to 15 degrees C. Our results thus show that the community structure of methanogens indeed affects the function of a cellulose-fermenting community with respect to temperature response.  相似文献   

14.
The interaction of insulin with phospholipids   总被引:1,自引:1,他引:0       下载免费PDF全文
1. A simple two-phase chloroform–aqueous buffer system was used to investigate the interaction of insulin with phospholipids and other amphipathic substances. 2. The distribution of 125I-labelled insulin in this system was determined after incubation at 37°C. Phosphatidic acid, dicetylphosphoric acid and, to a lesser extent, phosphatidylcholine and cetyltrimethylammonium bromide solubilized 125I-labelled insulin in the chloroform phase, indicating the formation of chloroform-soluble insulin–phospholipid or insulin–amphipath complexes. Phosphatidylethanolamine, sphingomyelin, cholesterol, stearylamine and Triton X-100 were without effect. 3. Formation of insulin–phospholipid complex was confirmed by paper chromatography. 4. The two-phase system was adapted to act as a simple functional system with which to investigate possible effects of insulin on the structural and functional properties of phospholipid micelles in chloroform, by using the distribution of [14C]glucose between the two phases as a monitor of phospholipid–insulin interactions. The ability of phospholipids to solubilize [14C]glucose in chloroform increased in the order phosphatidylcholine<sphingomyelin<phosphatidylethanolamine<phosphatidic acid. Insulin decreased the [14C]glucose solubilized by phosphatidylcholine, phosphatidylethanolamine and phosphatidic acid, but not by sphingomyelin. 5. The significance of these results and the molecular requirements for the formation of insulin–phospholipid complexes in chloroform are discussed.  相似文献   

15.
Soft and calcareous tissues embedded in polyester resin may be cut on a sledge microtome to produce thin sections of 3-4 β thickness. Fixed tissues, dehydrated in ethyl alcohol, cleared in methyl benzoate and chloroform, are taken into a wide-necked bottle containing equal parts of polyester resin and chloroform with 0.75% catalyst. The bottle kept in water bath at 37°C is connected to a vacuum pump. With the evaporation of the chloroform under reduced pressure (approximately 10 mm Hg) infiltration is complete. Tissues transferred into a blocking form containing pure polyester resin with 1.5% catalyst are polymerized at 37° C until blocks are firm (48 hr or more). Blocks are prepared with at least 5 mm margin of plastic surrounding the tissue. The edge of the block adjacent to the knife is then filed at an angle of 45° to the cutting movement. Sections are cut with a wide-backed biplanar knife having a cutting edge of 40-44° positioned at an angle of 30° to the plastic block. As the resin is permeable to most stains, staining is carried out through the plastic Sections carried through staining procedures in wire baskets are floated onto slides and mounted in polystyrene; the cover-glass is compressed with a spring-clamp. Microscopic examination shows no staining of plastic, minimal shrinkage and good cellular detail.  相似文献   

16.
The purpose of this study is to evaluate whether sample preservation can affect the yield of nucleic acid extracts from environmental samples. Storage of microbial samples was studied using three sediment types of varying carbon contents (10–57% carbon of dry weight). Four different storage solutions were tested at three temperatures. Freezing of samples at ?20 °C or ?80 °C, either without preservative or in phenol–chloroform solution, retained nucleic acid quantities very efficiently. Storage of samples in phenol–chloroform solution at +4 °C also gave good yields except for sediment with extremely high-carbon content. Ethanol and RNAlater® preservation decreased nucleic acid yields drastically at all temperatures. To study how sample preservation may affect the result of microbial community analysis, one type of sediment was selected for length heterogeneity-PCR analysis and PCR cloning of the 16S rRNA genes. Ethanol and RNAlater® preservation caused a slight bias towards certain microbial types in the community analyses shown by underrepresentation of Bacteroidetes, Betaproteobacteria and Gammaproteobacteria-affiliated peak sizes and overrepresentation of Actinobacteria, Chloroflexi and Alphaproteobacteria-affiliated peak sizes. Based on the results of this study, preservation in phenol–chloroform solution can be recommended as an alternative storage method when freezing is not possible such as during extended field sampling; however, ethanol and RNAlater® may cause serious problems when used as preservatives for environmental samples containing humic acids.  相似文献   

17.
Dechlorination of Chloroform by Methanosarcina Strains   总被引:9,自引:6,他引:3       下载免费PDF全文
Dehalogenation of carbon tetrachloride, chloroform, and bromoform in pure cultures of Methanosarcina sp. strain DCM and Methanosarcina mazei S6 was demonstrated. The initial dechlorination product of chloroform was methylene chloride (dichloromethane), which accumulated transiently to about 70% of the added chloroform; trace amounts of chloromethane were also detected. The amount of chloroform dechlorinated per mole of methane produced was approximately 10 times greater than the ratio observed previously for tetrachloroethene dechlorination by these strains. The production of 14CO2 from [14C]chloroform and the absence of 14CH4 imply that processes in addition to reductive dechlorination operate.  相似文献   

18.
Aims: To determine the optimal DNA extraction method for the detection of Coxiella burnetii including the small‐cell variant (SCV) by real‐time PCR (qPCR) in clinical samples. Methods and Results: A duplex qPCR detecting two Coxiella burnetii gene targets (com1 and IS1111a genes) was developed. Each target in this PCR had a sensitivity of one copy number per reaction. DNA extraction methods were compared on spiked negative samples and included a silica column kit, a chloroform separation prior to a silica column method and a chloroform/phenol separation and DNA precipitation method. Conclusions: The silica column extraction method was more efficient at recovering C. burnetii DNA, from large‐cell and small‐cell variants, than a chloroform or chloroform/phenol method. The silica column method was useful on spiked human samples including serum, buffy coat and bone marrow samples. Significance and impact of study: This study demonstrated that a simple column kit method is efficient to use for the detection of C. burnetii in clinical samples including the SCV.  相似文献   

19.
The effects of carbon tetrachloride, methylene chloride and chloroform on phosphorylation of PO was examined. The results of the dose response curve revealed that carbon tetrachloride (0.67%), methylene chloride (2%) and chloroform (1%) induced phosphorylation of PO by approximately 4, 6, and 12-fold, respectively. PO was found to be phosphorylated on the serine residue, and the phosphorylation of the serine residue was markedly increased when PO was phosphorylated in the presence of these compounds. Since tumor promoters, carbon tetrachloride and chloroform, have been shown to activate protein kinase C in platelets it is postulated that the increased phosphorylation of PO may result from the activation of myelin associated protein kinase C. The presence of phospholipid sensitive Ca2+-dependent protein kinase (protein kinase C) in purified nerve myelin was demonstrated by increased phosphorylation of PO in the presence of Ca2+ and phosphatidylserine.  相似文献   

20.
The half-height linewidth (v 1/2) of the 1H-NMR spectra of dipalmitoylphosphatidylcholine vesicles changes abruptly at the phase transition temperature. In the absence of inhalation anesthetics, proton signals from the choline head group (hydrophilic interface) and acyl-chain tails (lipid core) change at the same temperature of 39.6 degrees C. The present study compared the effect of four inhalation anesthetics, i.e., methoxyflurane, chloroform, halothane and enflurane, upon the ligand-induced phase transition of phosphatidylcholine vesicle membranes at 37 degrees C. The anesthetics showed differential action upon the phase transition of the phospholipid vesicle membranes between the lipid core and the hydrophilic interface. The concentrations of anesthetics which induced the phase transition of the lipid core were about 2-fold greater than those required for the phase transition of the interfacial choline head groups. From the area under the proton signals of inhalation anesthetics in the NMR spectra, the maximum solubilities of methoxyflurane, chloroform and halothane in 2H2O at 37 degrees C were determined to be 0.671 . 10(-4), 2.637 . 10(-4) and 1.398 . 10(-4) (expressed as mole fractions), or 3.35, 13.17 and 6.98 mmol/1000 g 2H2O, respectively. The solubilities of the anesthetic vapor in 2H2O expressed as mole fractions according to Henry's law ere 9.586 . 10(-4), 6.432 . 10(-4) and 2.311 10(-4)/atm (1.013 . 10(5) Pa) partial pressure, respectively. The presence of phospholipid vesicles in 2H2O increased the solubility of the inhalation anesthetics. From difference between solubility in 2H2O and a dipalmitoylphosphatidylcholine vesicle suspension, the partition coefficients of methoxyflurane, chloroform and halothane between the phospholipid vesicle membranes and 2H2O were estimated. These values, calculated from the mole fractions, were 3364, 1660 and 3850, respectively at 37 degrees C.  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