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 437 毫秒
1.
Teicoplanin (teic) from Actinoplanes teichomyceticus is a glycopeptide antibiotic used to treat many Gram-positive bacterial infections. Glycopeptide antibiotics inhibit bacterial growth by binding to carboxy-terminal d-Ala-d-Ala intermediates in the peptidoglycan of the cell wall of Gram-positive bacteria. In this paper we report the derivatization of magnetic microspheres with teic (teic-microspheres). Fluorescence-based techniques have been developed to analyze the binding properties of the microspheres to two d-Ala-d-Ala terminus peptides. The dissociation constant for the binding of carboxyfluorescein-labeled d-Ala-d-Ala-d-Ala to teic on microspheres was established via fluorimetry and flow cytometry and was determined to be 0.5 × 10−6 and 3.0 × 10−6 mol L−1, respectively. The feasibility of utilizing microparticles with fluorescence methods to detect low levels (the limit of bacterial detection was determined to be 30 colon-forming units; cfu) of Gram-positive bacteria has been demonstrated. A simple microfluidic experiment is reported to demonstrate the possibility of developing microsphere-based affinity assays to study peptide–antibiotic interaction.  相似文献   

2.
The effects of the spin coating process parameters on the thickness of the SiOx layer of the BOPP/SiOx composite film were investigated. When the concentration of tetraethoxysilane (TEOS) increased from 12.5 vol% to 55% vol%, the SiOx thickness increased from about 80 nm to 470 nm. In the sol time range of 1.5 h to 5 h the SiOx layer thickness reached a maximum at about 4 h and the change of the thickness roughly matched the change of the silica colloidal sphere sizes in sol. When the spin-coating speed of the dispensing stage increased from 450 r/min to 500 r/min, the SiOx layer thickness drastically decreased from about 1.67 μm to 400 nm. While the spin-coating speed of the thinning and drying stage went up to 1200 r/min, the SiOx layer thickness was in the range of 330 nm to 390 nm. It was also found that the SiOx layer thickness was almost increased linearly from about 500 nm to 1.02 μm with the ratio of the commercial silica colloidal to the TEOS from 0.2 to 1.0. The water contact angles decreased to about 23.0° for the BOPP/Si-Sol composite film with 1.67 μm SiOx layer and about 4.0° for the BOPP/mixing Si-Sol composite film with 1.02 μm SiOx layer. Compared to BOPP, the light transparency of the BOPP/Si-Sol composite films decreased by about 5.5% with the SiOx layer from about 80 nm to 1.67 μm and by 7.0% for the BOPP/mixing Si-Sol composite film with the SiOx layer from about 350 nm to 1.02 μm respectively.  相似文献   

3.
Mainly through the extensive investigations of Nakagawa and coworkers, nickel peroxide has emerged as a powerful oxidant for organic compounds. For example, it has been used for the oxidation of alcohols to aldehydes or carboxylic acids,2,3 allylic hydroxyl (selectively) to the carbonyl group,4 α-ketols to α-diketones,5 α-glycols, α-ketols, α-hydroxyl and α-keto acids to cleavage products,6 aldehydes to amides or nitriles in the presence of ammonia,7 phenols to quinones,8,9 certain Schiff bases to benzoxazoles,10 amines to azo compounds or nitriles,11,12 diarylamines to hydrazines,13 carbazoles to dimers and trimers,14 hydrazones to diazo compounds,15,16 1-aminobenzotriazole to benzyne,17 N-substituted hydroxylamines to azoxy compounds,18 phenylacetonitrile to dimeric products,19 a thiouracil to a uracil derivative,20 thiols to disulfides and sulfides to sulfones,21 N-substituted phenothiazines to sulfoxides and sulfones,21 haloforms to hexahaloethanes.23  相似文献   

4.
Direct electrochemical determination of arsenate (AsV) in neutral pH waters is considered impossible due to electro-inactivity of AsV. AsIII on the other hand is readily plated as As0 on a gold electrode and quantified by anodic stripping voltammetry (ASV). We found that the reduction of AsV to AsIII was mediated by elemental Mn on the electrode surface in a novel redox couple in which 2 electrons are exchanged causing the Mn to be oxidised to MnII. Advantage is taken of this redox couple to enable for the first time the electrochemical determination of AsV in natural waters of neutral pH including seawater by ASV using a manganese-coated gold microwire electrode. Thereto Mn is added to excess (∼1 μM Mn) to the water leading to a Mn coating during the deposition of As on the electrode at a deposition potential of −1.3 V. Deposition of As0 from dissolved AsV caused elemental Mn to be re-oxidised to MnII in a 1:1 molar ratio providing evidence for the reaction mechanism. The deposited AsV is subsequently quantified using an ASV scan. AsIII interferes and should be quantified separately at a more positive deposition potential of −0.9 V. Combined inorganic As is quantified after oxidation of AsIII to AsV using hypochlorite. The microwire electrode was vibrated during the deposition step to improve the sensitivity. The detection limit was 0.2 nM AsV using a deposition time of 180 s.  相似文献   

5.
The electron paramagnetic resonance (EPR) signal and the residual number (αt) of volatile molecules per phenyl group of poly(styrene-co-divinylbenzene) in samples that had been swelled to saturation in a dilute solution of a nitroxide spin-probe (TEMPO or 4-oxo-TEMPO), dissolved in a volatile liquid, were monitored simultaneously as the system containing excess liquid was allowed to evaporate to dryness. The results showed that the characteristic motionally narrowed three-line EPR spectrum began to change when αt became equal to αg (the number of sorbed molecules per phenyl group of polymer at liquid-saturation). The ratio of the intensity of the low-field and high-field hyperfine peaks relative to the middle peak decreased monotonically to an asymptotic limit that was attained when αt became equal to αg (the number of residual adsorbed molecules per phenyl group of polymer at completion of the transition from the rubbery state to the glassy state). The EPR hyperfine pattern, from which the rotational correlation times were estimated, changed most significantly as αt decreased from αG to αg while exhibiting inflections at about α′s and α′g the compositions that mark, respectively, incipient desorption of adsorbed molecules and incipient transition from the rubbery state to the glassy state. The pattern between these inflections points, however, varied with the affinity of the solvent for the polymer.  相似文献   

6.
SnO2 nanocrystalline material was prepared with a sol-gel process and thin films of the nanocrystalline SnO2 were coated on the surface of bent optical fiber cores for gas sensing. The UV/vis absorption spectrometry of the porous SnO2 coating on the surface of the bent optical fiber core exposed to reducing gases was investigated with a fiber optical spectrometric method. The SnO2 film causes optical absorption signal in UV region with peak absorption wavelength at around 320 nm when contacting H2-N2 samples at high temperatures. This SnO2 thin film does not respond to other reducing gases, such as CO, CH4 and other hydrocarbons, at high temperatures within the tested temperature range from 300 °C to 800 °C. The response of the sensing probe is fast (within seconds). Replenishing of the oxygen in tin oxide was demonstrated by switching the gas flow from H2-N2 mixture to pure nitrogen and compressed air. It takes about 20 min for the absorption signal to decrease to the baseline after the gas sample was switched to pure nitrogen, while the absorption signal decreased quickly (in 5 min) to the baseline after switching to compressed air. The adhesion of tin oxide thin films is found to be improved by pre-coating a thin layer of silica gel on the optical fiber. Adhesion increases due to increase interaction of optical fiber surface and the coated silica gel and tin oxide film. Optical absorption spectra of SnO2 coating doped with 5 wt% MoO3 were observed to change and red-shifted from 320 nm to 600 nm. SnO2 thin film promoted with 1 wt% Pt was found to be sensitive to CH4 containing gas.  相似文献   

7.
Radioactivity of coals and fly ashes   总被引:1,自引:0,他引:1  
The level and the behavior of the naturally occurring primordial radionuclides 238U, 226Ra, 210Pb, 232Th, 228Ra and 40K in coals and fly ashes are described. The activity concentrations of the examined coals and originated from coal mines in Greece ranged from 117 to 435 Bq·kg−1 for 238U, from 44 to 255 Bq·kg−1 for 226Ra, from 59 to 205 Bq·kg−1 for 210Pb, from 9 to 41 Bq·kg−1 for 228Ra and from 59 to 227 Bq·kg−1 for 40K. These levels are comparable to those appeared in coals of different countries worldwide. The activity concentrations of the examined fly ashes and produced in coal-fired power plants in Greece ranged from 263 to 950 Bq·kg−1 for 238U, from 142 to 605 Bq·kg−1 for 226Ra, from 133 to 428 Bq·kg−1 for 210Pb, from 27 to 68 Bq·kg−1 for 228Ra and from 204 to 382 Bq·kg−1 for 40K. The results showed that there is an enrichment of the radionuclides in fly ash relative to the input coal during the combustion process. The enrichment factors (EF) ranged from 0.60 to 0.76 for 238U, from 0.69 to 1.07 for 226Ra, from 0.57 to 0.75 for 210Pb, from 0.86 to 1.11 for 228Ra and from 0.95 to 1.10 for 40K.  相似文献   

8.
Infrared spectra of CO-treated platinum hydrosols subsequently treated with acetylene, hydrogen, and oxygen reveal that v(CO)ads decreases from 2070 cm−1 with increasing gas-treatment time. This has been attributed to a reduction in the coverage of adsorbed CO. In Pt sol/CO/C2H2 systems, v(CO)ads decreases to a limiting value of ca. 2060 cm−1 after exposure to acetylene. In the Pt sol/CO/H2 systems, v(CO)ads decreases to ca. 2050 cm−1 after exposure to hydrogen gas. The lower frequency in the Pt sol/CO/H2 system has been attributed to CO adsorption on more active metal sites formed from the reduction of surface platinum oxides. Exposure of the CO-treated platinum hydrosols to O2 gas was found to cause the eventual disappearance of the v(CO)ads band in infrared spectra, which was attributed to oxidation of adsorbed CO to CO2 by weakly bound surface layers of platinum oxides formed by the oxygen treatment.  相似文献   

9.
Our objective in this work is to gain insight into the contrasting carcinogenic activities of vinyl chloride (definitely carcinogenic) and trans-dichloroethylene (apparently inactive). The initial metabolic step for each molecule is believed to be epoxidation of the double bond, and there is evidence indicating that for vinyl chloride, this epoxide (chlorooxirane) is its ultimate (direct-acting) carcinogenic form. This article presents the findings of a computational study of the reactive properties of trans-dichlorooxirane (the epoxide of trans-dichloroethylene). An ab initio SCF -MO procedure was used to determine the energy requirements for stretching the C? O and C? Cl bonds (SN1 reactivity) and to study the epoxide's SN2 interactions with ammonia, taken as a model nucleophile. The starting points were the oxygen- and chlorine-protonated forms of the epoxide. The structure of the system was reoptimized at each step along the various reaction pathways. The results of this work are compared to an analogous earlier study of the reactive properties of chlorooxirane. The chlorineprotonated C? Cl bonds are found to have much lower energy barriers to stretching than do the oxygen-protonated C? O bonds. In the SN2 processes, intermediate complexes are formed with ammonia by both the oxygen- and the chlorine-protonated epoxides; the latter complexes are the more stable. Based on our results, we propose two mechanisms (one SN1 and the other SN2) whereby trans-dichlorooxirane can interact with N7 of guanine to produce an adduct analogous to one formed by chlorooxirane, which has been found to be the primary in vivo DNA alkylation product of vinyl chloride and to which has been attributed the carcinogenicity of the latter. Overall, trans-dichlorooxirane is found to be chemically more reactive than chlorooxirane; this may help to account for the much lesser carcinogenic and mutagenic activities of trans-dichloroethylene, since the epoxide may be reacting with other cellular nucleophiles before it reaches the key site(s) at which the carcinogenic or mutagenic interaction would occur. We also offer some speculations concerning other possible factors related to the differing carcinogenicities of vinyl chloride and trans-dichloroethylene, such as ease of epoxide formation and the likelihood of oxygen protonation.  相似文献   

10.
The well-known failure of the Hammett Ho function to describe the prototropic reaction isotherms of aromatic carboxamides can be corrected by taking account of the hydration requirements of the prototropic reactions of the carboxamides relative to the hydration requirements of the dissociations of the primary amines used as indicators to establish the Ho scale. This approach has been successfully applied to benzamide and the naphthamides. For up to a 4:1 molar ratio of water to sulfuric acid, the amide acidity scale, HA and the Ho scale are simply related by HA = Ho — 2 log aw , where aw is the activity of water.  相似文献   

11.
A systematic study was carried out to understand the effect of structural modification of Cnmim+ moiety of CnmimBr (n = 6, 8 or 10) on the electrochemical behavior of uranium. The cyclic voltammetric study of the above room-temperature ionic liquids (RTIL) media revealed that with increase in the chain length the electrochemical window extended more towards the negative potential. This resulted in the possibility of conversion of U(VI) to U(III) or even possibly to, U metal via U(IV) (as UO2) when n ≥ 10. The diffusion coefficient of U(VI) was found to decrease from n = 6 to 8 to 10 due to the increasing order of viscosity of the RTIL. As a consequence, the activation energy was found to follow reverse order i.e. E act(C6mimBr) < E act(C8mimBr) < E act(C10mimBr). The conversion of UO2 2+ to UO2 was found to be quasi reversible and also exothermic while the entropy was found to decrease due to the reduction reaction. An UV–Vis spectroscopic study was also carried out to understand the local environment around uranium in aqueous and RTIL media. Among several cationic and anionic species, the predominance of UO2Cl4 2? in 7 M HCl with D4h coordination symmetry was observed. The decrease in symmetric stretching frequency of UO2 2+ in RTILs in comparison with aqueous system indicates that the bond strength of UO2 2+ in aqueous is less than that of UO2 2+ in RTILs.  相似文献   

12.
Abstract

Recent measurements of the heat capacity at constant pressure Cp for lead from 300 to 850°K have shown that Cp for liquid lead decreases continuously from the melting point to 850°K. Using data in the literature of density and velocity of sound, the dilation correction has been applied to Cp to obtain the heat capacity at constant volume Cv for liquid lead. Application of the dilation correction to solid lead gives a Cv curve which uncreases only about one joule/gm-atom-°K from 300 to 600°K, whereas the Cv curve for liquid lead decreases almost 5 joules/gm-atom-°K from 600 to 850°K. A careful assessment of the uncertainty in the quantities used in the dilation correction leads to an uncertainty in Cv of ± 2.5% (about one joule/gm-atom-°K), and thus the decrease in Cv for liquid lead is quite real.  相似文献   

13.
The phase β-K0.25Pb0.75F1.75 previously found in the KF-PbF2 system appears to be metastable at low temperatures relative to a mixture of orthorhombic PbF2 and a new phase suspected to be KPbF3 II. KPbF3 II transforms to KPbF3 I at 298.5°C at atmospheric pressure. The KPbF3 II/I transition line rises with pressure, but the substance appears to reversibly disproportionate above ~360°C, 5 kbar, possibly to a mixture of PbF2 and K4PbF6. Instead of β-K0.25Pb0.75F1.75, a mixture with this composition yielded, in addition to weak heat events due to the KPbF3 II/I transition, strong heat events at 254.5°C and atmospheric pressure (thermal hysteresis ~13°C) which were ascribed to the PbF2 orthorhombic/cubic transition. This transition rises with pressure to 673°C at 37.8 kbar.  相似文献   

14.
As prerequisite for the investigation of arsenic-binding proteins in plants, the general influence of different extraction parameters on the binding behaviour of arsenic to the plant protein pool was investigated. The concentration of the extraction buffer affected the extraction yield both for proteins and for arsenic revealing an optimal buffer concentration of 5 mM Tris/HCl, pH 8. The addition of 1 or 2% (w/v) SDS to the extraction buffer produced a two- to threefold enhancement of the total protein extraction yield but strongly suppressed the simultaneous extraction of arsenic from 80 ± 8% extraction yield obtained without SDS to 48 ± 2% in presence of 2% (w/v) SDS. The arsenic binding capacity of the protein fraction obtained after extraction with Tris buffer and protein precipitation by trichloroacetic acid in acetone was estimated to be 1.4 ± 0.6% independently on the original spiking concentration of arsenic provided in the form of monomethylarsonate to the extracts. Due to the low total protein concentrations of the plant extracts that varied in the range from 75 to 412 μg mL−1 depending on the extraction parameters, high arsenic concentrations of 263-1001 mg (kg protein mass)−1 resulted for spiking concentrations of 10 mg As L−1. The optimized protein isolation procedure was applied to plants grown under arsenic exposure and revealed a similar arsenic binding capacity as for the spiked protein extracts.  相似文献   

15.
Ortho-phenylene-bridged phosphinoborane (2,6-Cl2Ph)2B-C6H4-PCy2 1 was synthesized in three steps from commercially available starting materials. 1 reacts with H2 or H2O under mild conditions to form corresponding zwitterionic phosphonium borates 1-H2 or 1-H2O . NMR studies revealed both reactions to be remarkably reversible. Thus, when exposed to H2, 1-H2O partially converts to 1-H2 even in the presence of multiple equivalents of water in the solution. The addition of parahydrogen to 1 leads to nuclear spin hyperpolarization both in dry and hydrous solvents, confirming the dissociation of 1-H2O to free 1 . These observations were supported by computational studies indicating that the formation of 1-H2 and 1-H2O from 1 are thermodynamically favored. Unexpectedly, 1-H2O can release molecular hydrogen to form phosphine oxide 1-O . Kinetic, mechanistic, and computational (DFT) studies were used to elucidate the unique “umpolung” water reduction mechanism.  相似文献   

16.
Reaction kinetics and proposed mechanism for the oxidation of propane over diluted Mo1–V0.3–Te0.23–Nb0.125–O x are described. The kinetic study allowed determination of the orders of propane disappearance, propene formation, CO x formation, and acids formation. The results show that selective oxidation of propane to propylene over this catalyst follows the Langmuir-Hinshelwood mechanism. Deep oxidation of propane to carbon dioxide is first order with respect to hydrocarbon, and partial order (0.21) with respect to oxygen. The selective oxidation of propane to acrylic acid is half order with respect to hydrocarbon and partial order (0.11) with respect to oxygen, while water does not participate directly in propane transformation. The result also shows that the overall reaction consists of three parallel process channels. One main sequence of consecutive reactions leads to the desired product.  相似文献   

17.
Mathematical consideration of the course of neutralization of a solution containing a moles of a strong base (e.g. sodium hydroxide) and a moles of a salt of a weak dibasic acid (e.g. sodium carbonate) with a strong acid (e.g. hydrochloric acid) shows that the first derivative of the titration curve should contain three maxima. The first, corresponding to neutralization of most of the hydroxide, is small (with height proportional to a12), displaced slightly from the equivalence point, and disappears for a less than 0.01 mol l-1. The second corresponds to the conversion of most of the CO32- to HCO3- and its height is almost independent of a. The third corresponds to the conversion of HCO3- to H2CO3, with height proportional to a12. The two minima are independent of a and of the dissociation constants of the weak acid. These conclusions were examined experimentally and were extended to the titration of hydroxide contaminated with a small amount of carbonate.  相似文献   

18.
Density functional theory with the B3LYP hybrid functional and 6–31G* basis set was used to study the geometric and electronic structure of H2C = CHR (R = H, CH3, C2H5, C3H7, C4H9, and C5H11) olefins, their carbocations formed in the addition of the proton to the olefins, R′-S-H aliphatic thiols (R′ = H, CH3, C2H5, and C3H7), the products of the addition of thiols to carbocations, and the final products of the addition of thiols to olefins. The proton affinity of the olefins and the products of the addition of thiols to olefins was calculated. The conclusion was drawn that the limiting stage in the nonradical addition of thiols to olefins catalyzed by acids was proton transfer from the protonated reaction product to the olefin. The theoretical results were compared with the experimental data on the electrophilic addition of polymercaptan to heptene-1.  相似文献   

19.
From the study of the CuCuBrCu2?εSC solid cell in the range of cubic digenite and “high temperature” hexagonal chalcocite we have deduced the laws of variation of the deviation from stoichiometry and the holes concentration with the equilibrium partial pressure of sulfur (δ and p are found to be proportional to p14S2). The electronic model corresponding to the formation of associations (V×CuVCu) in the presence of neutral vacancies V×Cu allows one to explain these laws. In the low temperature range (range of “low temperature” hexagonal chalcocite and orthorhombic chalcocite) the study of thermal variations of Hall coefficient permits us to propose the following models: (a) the deviation from stoichiometry of the “low temperature” hexagonal chalcocite is due to the simple ionized vacancies VCu; (b) the deviation from stoichiometry of the orthorhombic chalcocite is due to the vacancies VCu and to the associations (V×CuV×Cu).  相似文献   

20.
A multi-pumping flow system (MPFS) for the spectrophotometric determination of dissolved orthophosphate and dissolved organic phosphorus in wastewater samples is proposed. The determination of orthophosphate is based on the vanadomolybdate method. In-line ultraviolet photo-oxidation is employed to mineralise organic phosphorus to orthophosphate prior to detection. A solenoid valve allows the deviation of the flow towards the UV-lamp to carry out the determination of organic phosphorus.Calibration was found to be linear up to 20 mg P L−1, with a detection limit (3sb/S) of 0.08 mg P L−1, an injection throughput of 75 injections h−1 and a repeatability (R.S.D.) of 0.6% for the direct determination of orthophosphate. On the other hand, calibration graphs were linear up to 40 mg P L−1, with a detection limit (3sb/S) of 0.5 mg P L−1, an injection throughput of 11 injections h−1 and a repeatability (R.S.D.) inferior to 2.3% for the procedures involving UV photo-oxidation.  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