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 234 毫秒
1.
Efficient Uptake of Cesium Ions by Rhodococcus Cells   总被引:1,自引:0,他引:1  
Ivshina  I. B.  Peshkur  T. A.  Korobov  V. P. 《Microbiology》2002,71(3):357-361
Bacteria of the genus Rhodococcus were found to be able to accumulate cesium by means of active transport and nonspecific sorption on the cell surface structures. The maximum removal (up to 97%) of cesium from a medium supplemented with ammonium acetate was observed at 28°C, pH 7.8–8.6, and an equimolar content (0.2 mM) of potassium and cesium ions in the medium. The most active cesium-accumulating rhodococcal strains may be useful in biological treatment of industrial wastewaters contaminated with radionuclides.  相似文献   

2.
Experiments were conducted to study the desorption characteristics and plant-availability of phosphate sorbed by some important variable-charge minerals including kaolinite, goethite and amorphous Al oxide. Phosphate desorption from the complexes of goethite-P, kaolinite-P and Al oxide-P by equilibration with 0.02M KCl, resin or some commonly used chemical extractants was slow compared to desorption from a permanent-charge mineral (montmorillonite). However, rice plants were not observed under P deficiency in a pot trial with a phosphate-mineral complex as the only P source for both the permanent-charge mineral and the variable-charge minerals at either 50% or 100% sorption saturation with the exception of goethite-P at 50% saturation. In the exceptional goethite-P treatment, plant P concentration (1.0 g kg–1) was on the threshold of P deficiency. From 15% to 31% of the applied P was recovered by the plants within a growing period of three months, depending on sorption saturation and mineral type. Both the dry matter yield and P uptake decreased with decreasing sorption saturation for all the tested complexes except for Al oxide-P100 (100% saturation). In the case of Al oxide-P100, Al toxicity may have occurred, for poor root growth and high Al concentration in the plants were observed. The effect of sorption saturation on the yield and P uptake of plant was obvious for kaolinite and goethite but not very significant for montmorillonite. Based on the recovery of applied P, the plant-availability decreased in the following order: kaolinite-P100 > goethite-P100 > Al oxide-P50 > montmorillonite-P100 > montmorillonite-P50 > kaolinite-P50 > goethite-P50. Fractionation of the sorbed P before and after plant uptake showed that most of the P uptake originated from the resin-exchangeable P fraction in montmorillonite-P complex, but came mainly from NaOH-extractable fractions in goethite-P complex, whereas all the resin-P, NaHCO3-P and NaOH-P fractions in kaolinite- and amorphous Al oxide-P complex made a contribution to P uptake.  相似文献   

3.
Two-dimensional electrophoresis was carried out on fractions of the cyanobacteriumSynechocystis 6308 (ATCC 27150). Phycobilisomes isolated fromSynechocystis in 0.75M potassium phosphate buffer, pH 6.8, (KPi) plus Triton X-100 showed 5 prominent polypeptides when examined by two-dimensional electrophoresis. Ultracentrifugation of cells broken in KPi lacking Triton yielded three fractions, a membrane-containing pellet, a green supernatant, and a less dense yellow supernatant. The three fractions yielded a total of 272 polypeptides visualized by silver staining of two-dimensional gels. Fourteen polypeptides were found only in the yellow fraction, 14 polypeptides were found only in the green fraction, and 16 polypeptides were found only in the membrane fraction; 23 polypeptides were found in all three fractions. The crude Triton-containing KPi extract contained 62, and a 50 mM HEPES extract contained 55, of the 272 polypeptides visualized in these fractions. Two-dimensional electrophoresis combined with subcellular fractionation may be useful tools for examining changes in polypeptide composition caused by nitrogen starvation.  相似文献   

4.
Ruthenium red, a powerful inhibitor of Ca2+ transport by mitochondria, does not inhibit the active Ca2+ uptake by sarcoplasmic reticulum isolated from rabbit skeletal muscle promoted by 5 mM ATP-Mg in the presence or absence of potassium oxalate. Although concentrations of ruthenium red up to 100 μM do not affect the active uptake of Ca2+, 25 μM of the inorganic dye inhibit the passive binding of Ca2+ by about 50%. This inhibitory effect is observed in sarcoplasmic reticulum even after its lipid fraction is extracted with acetone.Although active Ca2+ uptake by sarcoplasmic reticulum is not inhibited by ruthenium red, in the absence of oxalate it inhibits significantly the Ca2+-dependent ATPase activity but not the Mg2+-ATPase. However, if potassium oxalate is present, the Ca2+-stimulated ATPase is not sensitive to the dye. It is not clear how oxalate functions to protect the Ca2+-ATPase against the inhibitor effect of ruthenium red.The high sensitivity to ruthenium red of the Ca2+ transport mechanism in mitochondria as compared to the Ca2+ transport in sarcoplasmic reticulum may be useful in determining the extent to which each organelle functions in the cell to regulate intracellular free Ca2+.  相似文献   

5.
Sorption isotherms of Norfloxacin (NOF) to different fractions from six typical sediments in China were determined to compare the NOF sorption behavior and contribution of different fractions to total sorption. All sorption isotherms were nonlinear and fitted well with the Freundlich model. Sorption coefficients (K f) by original sediments changed in larger magnitude, from 114 (mg/g)/(mg/L)n to 5271 (mg/g)/(mg/L)n, and black carbon with more aromatic carbon has more sorption capacity and nonlinearity. The sorption capacity K f values were found to significantly correlate with SSA (specific surface area), OC (organic carbon), BC (black carbon), and TON (total organic nitrogen) (p < 0.05), but had no obvious relation with pH, CEC (cation exchange capacity), TOC/TON, and BC/TOC. The DOC removed, NaOH extracted, and 375°C heated fractions showed more nonlinear sorption than the original sediments, suggesting more heterogeneous sorption sites in these fractions. Among different sediment fractions, the 375°C heating fractions were responsible for >50% of the total NOF sorption over the whole concentration range. The contribution of DOC removed fractions to the total sorption was the highest at higher NOF concentration.  相似文献   

6.
The crude methanolic extract of Andrachne cordifolia Muell. (Euphorbiaceae) and its various fractions in different solvent systems (chloroform, ethyl acetate and n- butanol) were screened for antibacterial and antifungal activities. Crude extract and subsequent fractions demonstrated moderate to excellent antibacterial activities against the tested pathogens. Highest antibacterial activity was displayed by both chloroform and ethyl acetate fractions (100%) followed by the crude extract (68%) against Salmonella typhi. Similarly, crude extract and its subsequent fractions showed mild to excellent activities in antifungal bioassay with maximum (76%) antifungal activity against Microsporum canis by the chloroform fraction followed by the crude extract (65%).  相似文献   

7.
Because of low net production in arctic and subarctic surface water, dissolved organic matter (DOM) discharged from terrestrial settings plays an important role for carbon and nitrogen dynamics in arctic aquatic systems. Sorption, typically controlling the export of DOM from soil, may be influenced by the permafrost regime. To confirm the potential sorptive control on the release of DOM from permafrost soils in central northern Siberia, we examined the sorption of DOM by mineral soils of Gelisols and Inceptisols with varying depth of the active layer. Water‐soluble organic matter in the O horizons of the Gelisols was less (338 and 407 mg C kg?1) and comprised more dissolved organic carbon (DOC) in the hydrophobic fraction (HoDOC) (63% and 70%) than in the O horizons of the Inceptisols (686 and 706 mg C kg?1, 45% and 48% HoDOC). All A and B horizons from Gelisols sorbed DOC strongly, with a preference for HoDOC. Almost all horizons of the Inceptisols showed a weaker sorption of DOC than those of the Gelisols. The C horizons of the Inceptisols, having a weak overall DOC sorption, sorbed C in the hydrophilic fraction (HiDOC) stronger than HoDOC. The reason for the poor overall sorption and also the preferential sorption of HiDOC is likely the high pH (pH>7.0) of the C horizons and the smaller concentrations of iron oxides. For all soils, the sorption of HoDOC related positively to oxalate‐ and dithionite–citrate‐extractable iron. The A horizons released large amounts of DOC with 46–80% of HiDOC. The released DOC was significantly (r=0.78, P<0.05) correlated with the contents of soil organic carbon. From these results, we assume that large concentrations of DOM comprising large shares of HiDOC can pass mineral soils where the active layer is thin (i.e. in Gelisols), and enter streams. Soils with deep active layer (i.e. Inceptisols), may release little DOM because of more frequent infiltration of DOM into their thick mineral horizons despite their smaller contents of reactive, poorly crystalline minerals. The results obtained for the Inceptisols are in agreement with the situation observed for streams connecting to Yenisei at lower latitudes than 65°50′ with continuous to discontinuous permafrost. The smaller sorption of DOM by the Gelisols is in agreement with the larger DOM concentrations in more northern catchments. However, the Gelisols preferentially retained the HoDOC which dominates the DOC in streams towards north. This discrepancy can be explained by additional seepage water from the organic horizons that is discharged into streams without intensive contact with the mineral soil.  相似文献   

8.
To establish relationships between soil phosphorus (P) fractions and leaf P, a mycorrhizal species (Plantago lanceolata L.) was compared with a typically non-mycorrhizal species (Rumex acetosella L.) in a glasshouse experiment. The plants were grown in 40 soils from non-fertilised, abandoned pastures or abandoned arable fields and leaf P concentration were found to be related to various soil P fractions after six weeks of growth. The differences in the P fractions in soil can account for a large share of the variation in leaf P concentration in both species, but the two species differed in their utilisation of P fractions. Leaf P concentration of R. acetosella was more related to extractable soil P than that of P. lanceolata. Rumex acetosella showed a higher maximum P concentration. The P fractions accounting for the largest share of the variation in leaf P concentration was the Bray 1 extractable and the weak oxalate (1 mM) extractable P, and for P. lanceolata also the Na2SO4+NaF extractable P fraction. P extracted with these methods accounted for up to 80% of the variation in P concentration in leaves of R. acetosella and 65% of the variation in leaves of P. lanceolata. More P extractable with weak oxalate, Na2SO4+NaF and strong oxalate (50 mM) was released from the soil than was taken up by the plants during the experimental period. The Bray 1 extractable P fraction, however, decreased in both unplanted and planted soils. Phosphatase release was not induced in any of the plants during the experimental period, indicating that they were not mobilising soil organic P. However, some of the methods extracted a large share of the organic P and still explained much of the variation in leaf P concentration. Mycorrhizal colonisation of P. lanceolata was inversely related to the extractable soil P. The consistently fast P uptake of R. acetosella indicates that this species have a high demand for P. The differences in P utilisation between R. acetosella and P. lanceolata could be caused by their different mycorrhizal status.  相似文献   

9.
A Perigo-type antibacterial factor (PTF) was produced when tryptone (a pancreatic digest of casein) medium was heated with nitrite at 121°C for 20 min. This PTF was inhibitory against Staphylococcus aureus, Bacillus subtilis and Clostridium botulinum, but was not against Escherichia coli and Salmonella typhimurium. The inhibitory activity varied with the concentration of nitrite (5 ~ 100 ppm) and tryptone (1, 2, 4%), and with pH (4, 5, 6, 7). The maximum inhibitory activity was observed when the medium containing 4% tryptone and 0.2% thioglycolate was heated with more than 50 ppm nitrite at pH 6. The tryptone was separated into three fractions by gel filtration and PTF was produced in every fraction, although the inhibitory activity was different in each. Our PTF might be unstable towards oxygen because its activity was lost completely by shaking for more than 16 hr.  相似文献   

10.
Sorption kinetics and isotherms of phenol by four carbonaceous sorbents (activated carbon (AC), mesoporous carbon (MPC), bamboo biochar (BBC) and oak wood biochar (OBC)) were compared in this study. MPC has the fastest sorption rate and initial sorption potential, which were indicated by sorption rate constants and initial sorption rate “h” in a pseudo-second-order kinetic model. The ordered and straight pore structure of MPC facilitated the accessibility of phenol. The AC showed the greatest sorption capacity towards phenol with maximum sorption of 123 mg/g as calculated by the Langmuir model. High surface area, complexity of pore structure, and the strong binding force of the ππ electron-donor-acceptor interaction between phenol molecules and AC were the main mechanisms. The BBC and OBC had much slower sorption and lower sorption capacity (33.04 and 29.86 mg/g, respectively), compared to MPC (73.00 mg/g) and AC, indicating an ineffective potential for phenol removal from water.  相似文献   

11.
The leaves of Costus pictus are sour in taste due to the presence of oxalic acid in the leaves. Different stages of leaves were collected and the samples were designated as stage one, stage two and stage three. It was found that oxalate content and oxalate oxidase activity were maximum in second leaf stage followed by first leaf stage and third leaf stage. Drying causes substantial loss of oxalate content and complete loss of oxalate oxidase activity. With various solvents water recovered more oxalate followed by methanol and ethanol while oxalate oxidase activity was maximum in ethanol followed by methanol and water. The ethanol or methanol extract of second leaf stage of C. pictus can be used for isolating active principles. The oxalate oxidase from C. pictus can be used as a cheap source of oxalate oxidase enzyme which is used in oxalate determination in biological fluids. Moreover, the sensitivity of oxalate determination employing oxalate oxidase from C. pictus will be more as oxalate oxidase in C. pictus has K m 20 times lesser than the oxalate oxidase enzyme from barley seedling.  相似文献   

12.
A sequential extraction scheme was combined with sorption isotherm analysis in order to investigate sorption of sewage sludge-derived Cu and Zn to the A-horizon of a humic-gley soil as a whole, and to the operationally defined exchangeable (1?M MgCl2), carbonate (1?M NaOAc), Fe/Mn oxide (0.04?M NH2OH.HCl), and organic (0.02?M HNO3+30% H2O2) soil fractions. Sorption parameters were compared for a sample of sludge leachate (with 97.4% of Cu and 63.2% of Zn present as dissolved metal-organic matter complexes, as calculated by geochemical modeling involving MINTEQA2 and verified using an ion exchange resin method) with that of a reference solution exhibiting the same chemical characteristics as the leachate, except for the presence of dissolved organic material. Dissolved metal-organic matter complexes were found to significantly (P<0.05) depress sorption to the bulk soil and each fraction. The greatest depression of Cu and Zn sorption was observed for the exchangeable, carbonate, and Fe/Mn oxide fractions, while the organic fraction of the soil was the least affected. This reflects a greater affinity for the exchangeable, carbonate, and Fe/Mn oxide fractions by the free divalent metal (Cu2+, Zn2+), with sorption by these fractions attributed to cation exchange, chemisorption, and co-precipitation processes. The sorption characteristics of the organic fraction indicated that Cu and Zn sorption by soil organic matter mostly involved dissolved metal-organic matter complexes. This may be attributed to hydrophobic interactions between nonpolar regions of the dissolved metal-organic matter complexes and solid-phase soil organic matter.  相似文献   

13.
The crude methanolic extract and subsequent fractions of Teucrium royleanum (Labiatea) were screened for antibacterial and antifungal activities. Against tested pathogens, crude extract and subsequent fractions demonstrated moderate to excellent antibacterial activities. Highest antibacterial activity was displayed by the ethyl acetate fraction against S. typhi (100%), against E.coli (76.7%) and against P. aerugenosa (70.8%) followed by the chloroform fraction against S. typhi (85.7%). Similarly, the crude extract and its subsequent fractions showed mild to excellent activities in the antifungal bioassay with maximum antifungal activity against M. canis (87%) by the chloroform fraction followed by the ethyl acetate (71%) and n-butanol (70%) fractions.  相似文献   

14.
The microsomal fraction isolated from sea urchin H. pulcherrimus eggs has the ability to actively accumulate Ca2+ in the presence of ATP. The Ca2+ uptake was sustained by addition of oxalate and was apparently insensitive to sodium azide. The sequestered microsomal Ca was readily released by the divalent cation ionophore A23187. The microsomal fraction obtained from fertilized eggs accumulated Ca2+ about five times more quickly than did that from unfertilized eggs. The increased Ca2+ uptake by microsomal fraction obtained from fertilized eggs was due to an increase in the maximum velocity of Ca2+ uptake and there was no difference in Km for calcium between the two fractions.  相似文献   

15.

Antimicrobial peptides (AMPs) are molecules present in several life forms, possess broad-spectrum of inhibitory activity against pathogenic microorganisms, and are a promising alternative to combat the multidrug resistant pathogens. The aim of this work was to identify and characterize AMPs from Capsicum chinense fruits and to evaluate their inhibitory activities against yeasts of the genus Candida and α-amylases. Initially, after protein extraction from fruits, the extract was submitted to anion exchange chromatography resulting two fractions. Fraction D1 was further fractionated by molecular exclusion chromatography, and three fractions were obtained. These fractions showed low molecular mass peptides, and in fraction F3, only two protein bands of approximately 6.5 kDa were observed. Through mass spectrometry, we identified that the lowest molecular mass protein band of fraction F3 showed similarity with AMPs from plant defensin family. We named this peptide CcDef3 (Capsicum chinense defensin 3). The antifungal activity of these fractions was analyzed against yeasts of the genus Candida. At 200 μg/mL, fraction F1 inhibited the growth of C. tropicalis by 26%, fraction F2 inhibited 35% of the growth of C. buinensis, and fraction F3 inhibited all tested yeasts, exhibiting greater inhibition activity on the growth of the yeast C. albicans (86%) followed by C. buinensis (69%) and C. tropicalis (21%). Fractions F1 and F2 promoted membrane permeabilization of all tested yeasts and increased the endogenous induction of reactive oxygen species (ROS) in C. buinensis and C. tropicalis, respectively. We also observed that fraction F3 at a concentration of 50 µg/mL inhibited the α-amylase activities of Tenebrio molitor larvae by 96% and human salivary by 100%. Thus, our results show that fraction F3, which contains CcDef3, is a very promising protein fraction because it has antifungal potential and is able to inhibit the activity of different α-amylase enzymes.

  相似文献   

16.
The sulphur nutrition of three isolates ofAlternaria tenuis Auct., isolated from the diseased leaves ofMangifera indica L.,Musa paradisiaca L. andPsidium guajava L., was studied. They were grown on the medium devoid of sulphur as well as on media containing various sources of sulphur viz., ammonium sulphate, sodium hyposulphite, sodium thiosulphate, magnesium sulphate, potassium sulphate, potassium metabisulphite, zinc sulphate and thiourea. Sodium hyposulphite, sodium thiosulphate, magnesium sulphate, potassium sulphate and zinc sulphate were generally found to be satisfactory sources for the growth of all the isolates under study. Poor growth of the different isolates was observed on the medium devoid of sulphur.  相似文献   

17.
Summary A fraction of erythrocyte Band 3 (M r , 93,000) glycoprotein that demonstrates decreased autophosphorylation in membranes from myotonic muscular dystrophy patients is demonstrated. Sequential affinity chromatography of Triton X-100 solubilized erythrocyte membrane proteins separated three specifically retained glycoprotein fractions on a Ricin Communis I-Sepharose 4B column. One fraction contains a portion of the major sialoglycoprotein (apparentM r , 78,000) and is specifically eluted from the column by 10mm NaCl and 100mm d-galactose (10/100). The two other glycoprotein fractions are eluted by 100mm NaCl, 10mm d-galactose (100/10) and 100mm NaCl, 100mm d-galactose (100/100). The composition of both fractions contains greater than 95% Band 3 (apparentM r , 93,000) glycoprotein.The quantities of glycoprotein in each fraction obtained from erythrocytes of myotonic dystrophy patients did not differ from the quantities obtained from control erythrocytes. Following endogenous protein kinase incubations of ghosts with [-32P]ATP, the specific [32P] phosphorylation of the 10/100 and 100/10 fractions are identical. The 100/100 fraction, which makes up approximately 3% of the total erythrocyte membrane protein, demonstrates a different pattern for myotonic dystrophy patients; specific phosphorylation was reduced by 50% relative to activity in control experiments. These findings are consistent with previous experiments that demonstrated decreased autophosphorylation of the glycoprotein portion of Band 3 (Roses & Appel, 1975,J. Membrane Biol. 20: 51) and are consistent with the autosomal dominant mode of inheritance in this disease.  相似文献   

18.
Sisler EC 《Plant physiology》1980,66(3):404-406
An ethylene binding component(s) has been partially purified from mung bean sprouts. Tissue was homogenized in 0.3 molar sucrose and 0.2 molar potassium phosphate buffer (pH 7.0). The homogenate was centrifuged, and resuspended fractions were assayed by incorporating them onto cellulose fibers (0.7 grams per milliliter). These were exposed to [14C]ethylene (3.7 × 10−2 microliters per liter of 120 millicurie per millimole) in the presence or absence of 1000 microliters per liter unlabeled ethylene. The cellulose was transferred to separate containers and the [14C]ethylene was absorbed in mercury perchlorate and counted. Distribution of ethylene binding to various fractions was: 0 to 3,000g, 3%; 3,000 to 12,000g; 4%; 12,000 to 100,000g, 69%; cellular debris, 24%; 100,000g supernatant, 0%. Adjustment of the pH to 4.0 precipitates the ethylene-binding component. Neutralization, addition of Triton X-100, and readjustment of the pH to 4.0 “solubilized” most of the binding component. Further purification was obtained by chromatography on CM-Sephadex in 10 millimolar potassium acetate buffer, (pH 5.0) containing 1% Triton X-100. Elution was with 200 millimolar potassium phosphate (pH 6.0) containing 1% Triton X-100. Upon treatment of the Triton “solubilized” component with cold acetone, over 90% of the binding capacity was lost. Extraction of the acetone-precipitated residue with 2% Triton X-100 restored some of the binding capacity which was found in the soluble fraction. The pH optimum for binding is 6.0. Passing the Triton X-100 extract of the acetone powder through Sepharose 6B provides considerable purification. The binding component moved ahead of most of the protein.  相似文献   

19.
Homogenates of dormant cysts of Artemia salina were fractionated by differential centrifugation. RNA was prepared from the various fractions and tested for stimulatory activity in a [14C]leucine incorporating Escherichia coli system. The highest specific activity was found in the RNA extracted from a cytoplasmic fraction sedimenting at 15,000 g. Some activity was associated with the soluble and crude ribosomal fractions, while the RNA extracted from the crude nuclear fraction was less active.The 15,000 g sediment was purified by centrifugation in a sucrose density gradient. The active material formed a characteristic, colored band at a buoyant density of about 1.17 g/ml. The banding fraction was mainly composed of endoplasmic vesicles and mitochondria. The specific activity of the extracted RNA was further increased when the 15,000 g sediment was treated with buffered 20–100 mM EDTA (with or without 0.1% Triton X-100) before banding.Sedimentation analysis of the active RNA from the purified 15,000 g fractions revealed three distinct absorption peaks at 28 S, 18 S, and 16 S, apparently representing cytoplasmic and mitochondrial rRNA. The 28 S and 18 S peaks were reduced by EDTA treatment, but only to a certain limit. By gel electrophoresis a number of additional components were resolved, including 4 S and 5 S RNA. The template activity showed a heterodisperse distribution with a maximum at 17–20 S, not correlated with the 16 S peak. Isolated 18 S and 28 S rRNA had very low activity.The experiments suggest that in Artemia cysts an appreciable amount of messengerlike RNA is associated with mitochondria and/or endoplasmic vesicles carrying ribosomal monomers.  相似文献   

20.
The subcellular distributions of potassium, sodium, magnesium, calcium and chloride have been determined for rabbit cerebral cortex. After homogenization and differential centrifugation, the following percentages of ions were associated with the particulate fraction (nuclear, mitochondrial, synaptic vesicles, and microsomal): (a) 19% of the total potassium; (b) 22% of the total sodium; (c) 77% of the total calcium; (d) 69% of the total magnesium; and (e) less than 2% of the total chloride. However, the sum of the potassium and sodium content in each of the particulate fractions was greater than the sum of the calcium and magnesium content. After hypo-osmotic shock of the crude mitochondrial fraction (MT), more sodium than potassium (μmol/g wet wt.) was associated with the mitochondrial (M1) and synaptic vesicle (M2) fractions. The molar ratio of sodium to potassium was 1·4 for M1 and 4·5 for M2. The association of 22Na+ with the particulate fractions was further studied by the method of equilibrium dialysis. The data from both types of experiments indicate that a large fraction of the sodium in cortical tissue appears to be in a bound state.  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