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 15 毫秒
1.
Various amino acid derivatives of monascus pigments were synthesized. The effects of pigment derivatives on the pigment adsorption ratio, electrophoretic mobility (EPM) of bacterial cells, and antibacterial activity were investigated under varying conditions of pigment type, pigment concentration, pH, and ionic strength. Two hydrophobic and two hydrophilic derivatives were selected as model pigments. There was a close relationship between the antimicrobial activity and the pigment adsorption ratio. Against Escherichia coli, the hydrophobic l-Tyr and l-Phe derivatives (log P = 3.18 and 3.57) exhibited high antimicrobial activities (MIC = 8 and 16 mg/L) and high cellular adsorption ratios (9.6 and 10.9 mg/L). The hydrophilic l-Glu and l-Asn derivatives (log P = 1.40 and 0.47) exhibited low activities (MIC = 64 and 128 mg/L) and low adsorption ratios (4.7 and 4.0 mg/L). The electrophoretic mobility of 11 different bacteria varied between −1.93 × 10−8 and −1.19 × 10−8 m2 V−1 s−1 regardless of Gram+ or Gram. The l-Phe derivative showed low MIC values (high antimicrobial activities) against bacteria with a high electrophoretic mobility. A positive linearity between the pigment adsorption ratio and the electrophoretic mobility was established. When the four pigment derivatives were added to E. coli solutions, the electrophoretic mobility of cells in all cases sharply increased with an increasing pigment concentration. The mobility value was high for hydrophobic pigment derivatives in descending order of l-Phe (0.8 × 10−8 m2 V−1 s−1), l-Tyr (0.68 × 10−8 m2 V−1 s−1), l-Glu (0.46 × 10−8 m2 V−1 s−1), and l-Asn (0.44 × 10−8 m2 V−1 s−1). Additional adsorption of the hydrophobic derivatives probably occurred due to a hydrophobic interaction between the pigment and the pigment-coated cells. The electrophoretic mobility decreased gradually with an increasing pH and/or ionic strength with both addition and no addition of the pigment derivatives. The pattern of change of the pigment adsorption ratio under varying pH and/or ionic strength values was similar to the pattern for electrophoretic mobility.  相似文献   

2.
Column technology for capillary electrochromatography   总被引:4,自引:0,他引:4  
Column technologies for capillary electrochromatography (CEC) are reviewed. To achieve high efficiency, the inner diameters of open-tubular and packed columns should be less than 25 and 200 μm, respectively. To obtain acceptable separation speed under typical CEC conditions (e.g. 30 kV, 1 mm s−1 electroosmotic flow velocity, and 2–4×10−8 m2 V−1 s−1 electroosmotic mobility) the column lengths for open-tubular and packed columns should be less than 120 and 60 cm, respectively. Capillary CEC columns are generally classified into three types: packed, open-tubular, and continuous-bed or monolithic. The various column preparation procedures and the advantages and disadvantages of each column type are discussed in detail.  相似文献   

3.
Mealor D  Townshend A 《Talanta》1968,15(12):1477-1480
Methods are described for the determination of cyanide (10−8–10−5M and sulphide (10−7–10−5 M) based on the de-inhibitory effect of these ions on invertase inhibited by mercury(II) or by silver. Iodine (0.1–3 μg) may be determined by its inhibition of invertase.  相似文献   

4.
Mealor D  Townshend A 《Talanta》1968,15(12):1371-1376
Thiourea has been found to enhance the inhibition of invertase by silver ions. The effect is applied to the determination of 1–5 × 10−7 M silver and of 10−7–10−8 M thiourea. A mechanism is suggested for the enhancement.  相似文献   

5.
Cha KW  Park CI  Park SH 《Talanta》2000,52(6):689-989
Uranium(VI) complexed with aluminon (3-[bis(3-carboxy-4-hydroxy-phenyl)methylene]-6-oxo-1,4-cyclohexadiene-1-carboxylic acid triammonium salt) was determined by adsorptive cathodic stripping voltammetry (ACSV) using a hanging mercury drop electrode. Trace uranium(VI) and zinc(II) can be simultaneously determined in a single scan in the presence of aluminon and urea. Optimal conditions were found to be: accumulation time; 180–200 s, accumulation potential; 50 mV versus Ag/AgCl, scan rate; 40 mV s−1, supporting electrolyte; 0.1 M sodium acetate buffer at pH 6.5–7.0, and concentration of aluminon; 1×10−6 M. The linear range of uranium(VI) and zinc(II) were observed over the concentration range 2–33 and 30–120 ng ml−1, respectively. The detection limit (S/N=3) are 0.2 ng ml−1 (uranium) and 30 ng ml−1 (zinc). A good reproducibility shows RSDs of 2.5–4.0% (n=10). The procedure offers high selectivity, with the presence of urea masking some metal ions.  相似文献   

6.
The rate coefficients of the reactions: (1) CN + H2CO → products and (2) NCO + H2CO → products in the temperature range 294–769 K have been determined by means of the laser photolysis-laser induced fluorescence technique. Our measurements show that reaction (1) is rapid: k1(294 K) = (1.64 ± 0.25) x 10−11 cm3 molecule−1 s−1; the Arrhenius relation was determined as k1 = (6.7 ± 1.0) x 10−11 exp[(−412 ± 20)/T] cm3 molecule−1 s−1. Reaction (2) is approximately a tenth as rapid as reaction (1) and the temperature dependence of k2 does not conform to the Arrhenius form: k2 = 4.62 x 10−17T1.71 exp(198/T) cm3 molecule−1 s−1. Our values are in reasonable agreement with the only reported measurement of k1; the rate coefficients for reaction (2) have not been previously reported.  相似文献   

7.
Rate constants for the reactions of OH with CH3CN, CH3CH2CN and CH2=CH-CN have been measured to be 5.86 × 10−13 exp(−1500 ± 250 cal mole−1/RT), 2.69 × 10−13 exp(−1590 ± 350 cal mole−1/RT and 4.04 × 10−12 cm3 molecule−1 s−1, respectively in the temperature range 298–424 K. These results are discussed in terms of the atmospheric lifetimes of nitrfles.  相似文献   

8.
Antigen I/II can be found on streptococcal cell surfaces and is involved in their interaction with salivary proteins. In this paper, we determine the adsorption enthalpies of salivary proteins to Streptococcus mutans LT11 and S. mutans IB03987 with and without antigen I/II, respectively, using isothermal titration calorimetry. In addition, protein adsorption to the cell surfaces was determined spectrophotometrically. S. mutans LT11 with antigen I/II, yielded a much higher, exothermic adsorption enthalpy at pH 6.8 (ranging from −2073 × 10−9 to −31707 × 10−9 μJ per bacterium) when mixed with saliva than did S. mutans IB03987 (−165 × 10−9 to −1107 × 10−9 μJ per bacterium) at all bacterial concentrations studied (5 × 109, 5 × 108, and 5 × 107 ml−1), largest effects per bacterium being observed for the lowest concentration. However, the enthalpy of salivary protein adsorption to S. mutans LT11 became smaller at pH 5.8. Adsorption isotherms for the S. mutans LT11 showed considerable protein adsorption at pH 6.8 (1.2–2.1 mg/m2), that decreased only slightly at pH 5.8 (1.1–1.6 mg/m2), with the largest amount adsorbed at the lowest bacterial concentration. This suggests that the protein(s) in the saliva with the strongest affinity for antigen I/II is (are) readily depleted from saliva. In conclusion, antigen I/II surface proteins on S. mutans play a determinant role in adsorption of salivary proteins through the creation of enthalpically favorable adsorption sites.  相似文献   

9.
Depending on the sulfur species, picomoles of different inorganic sulfur compounds can be detected and separated by HPLC in one arrangement in a sample volume less than 50 μl. The combination of fluorescence labelling of reduced inorganic sulfur compounds such as sulfide (S2−), sulfite(SO32− and thiosulfate (S2O32−) with monobromobimane followed by an extraction of elemental sulfur (S°) by chloroform treatment enables the detection of all mentioned sulfur compounds as well as sulfate (remaining aqueous phase) in the same sample. While the derivatized sulfur compounds could be detected by their fluorescence emission at 480 nm, elemental sulfur is identified by its UV absorption at 263 nm. Sulfate in the remaining aqueous phase is detected by HPLC with indirect UV detection at 254 nm. Detection ranges for the different sulfur compounds examined are as follows: sulfide (5 μM to 1.5 mM), sulfite (5 μM to 1.0 mM), thiosulfate (1 μM to 1.5 mM), elemental sulfur (2 μM to 32 mM) and sulfate (5 μM to >1 mM).  相似文献   

10.
A number of experimental parameters have been optimized for the separation of 26 metal ions, including alkali, alkaline earth, transition and lanthanide metal ions. Experimental parameters that were evaluated included nature of indirect-detection reagent, pH of electrolyte, concentration of complexing agent and nature of the surface of the capillary; unbonded and C1 and C18 bonded phases were studied. In addition the effect of internal diameter on linearity and signal-to-noise ratio was examined, and separation efficiency was determined for a variety of experimental conditions. Detection limits (signal-to-noise RATIO = 3) were ca. 1 μg/ml for the lanthanides, ca. 0.6 μg/ml for transition and alkaline earth ions and ca. 0.1–0.8 μg/ml for alkali metal ions. The average relative standard deviations of were 3.7, 5.1 and 2.5% on unbonded, C1 and C18 capillaries, respectively. Whereas conventional regression analysis suggested that the calibration curves were linear over the range of 1·10−5 to 4·10−4 mol/l, sensitivity plots showed that the results were actually linear to within 6% only over the range of 2.5·10−5 to 4·10−4 mol/l.  相似文献   

11.
Trace amounts of nickel(II) can function as a trigger (=reaction initiator) in an autocatalytic reaction with the sodium sulfite/hydrogen peroxide system. Based on this finding, sub-μg L−1 levels of nickel(II) were determined by a time measurement using the autocatalytic reaction. The detection range using the above method was 10−9–10−5 M, the detection limit (3σ) was 8.1 × 10−10 M (0.047 μg L−1), and the relative standard deviation was 2.66% at nickel(II) concentration of 10−7 M (n = 7). This method was applied to length detection-flow injection analysis. The detection range for the flow injection analysis was 2 × 10−9–2 × 10−3 M. The detection limit (3σ) was 1.4 × 10−9 M (0.082 μg L−1), and the relative standard deviation was 1.86 at initial nickel(II) concentration of 10−6 M (n = 7).  相似文献   

12.
Sultan SM  Hassan YA  Abulkibash AM 《Talanta》2003,59(6):1073-1080
For the first time, promethazine hydrochloride chemiluminescence emission was monitored. The paper describes a new, specific and highly sensitive flow injection (FI) method for the determination of promethazine hydrochloride using both a peristaltic and a syringe pump. The method was based on the chemiluminescence emission intensity produced as a result of its oxidation reaction with permanganate in sulfuric acid medium. Reaction variables were thoroughly investigated employing chemometrical methods with few number of experiments. The optimum system and chemical conditions were 2.1519×10−4 mol l−1 permanganate in 0.01 mol l−1 sulfuric acid when operating the peristaltic pump at a flow rate of 45 μl s−1 and injecting the drug by a syringe pump operated at a speed of 40 μl s−1. The method was found to be applicable in the concentration range of promethazine hydrochloride between 1.558×10−5 and 1.8697×10−3 mol l−1 with a linear calibration plot of 0.992 correlation coefficient and the following equation: I=92.74+0.08048C. The method adopted proved to be highly suitable for the assay of promethazine hydrochloride in drug formulations without fear of interferences in dosage form.  相似文献   

13.
Inam R  Somer G 《Talanta》1998,46(6):1347-1355
The polarographic reduction of lead in the presence of selenite gives rise to an additional peak corresponding to the reduction of lead (Pb) on adsorbed selenium (Se) on mercury at −0.33 V. The selenium and lead content can be determined using this peak by the addition of a known amount of one of these ions first and then the second ion. The linear domain range of lead is 5.0×10−7–2.0×10−5 M and for selenium 5.0×10−7–1.0×10−5 M. Using this method 4.90×10−7 M Se(IV) and 1.47×10−6 M Pb(II) in a synthetic sample could be determined with a relative error of +2.0% and 1.8%, respectively (n=4). A recovery test after acid digestion for a synthetic sample was 97% for selenium and 96.5% for lead. The method was applied to 1 ml of digested blood, and 328±23 μg l−1 Se(IV) and 850±62 μg l−1 Pb(II) could be determined with a 90% (n=5) confidence interval.  相似文献   

14.
Wang Q  Li N 《Talanta》2001,55(6):243-1225
The thiolactic acid (TLA) self-assembled monolayer modified gold electrode (TLA/Au) is demonstrated to catalyze the electrochemical response of norepinephrine (NE) by cyclic voltammetry. A pair of well-defined redox waves were obtained and the calculated standard rate constant (ks) is 5.11×10−3 cm s−1 at the self-assembled electrode. The electrode reaction is a pseudo-reversible process. The peak current and the concentration of NE are a linear relationship in the range of 4.0×10−5–2.0×10−3 mol l−1. The detection limit is 2.0×10−6 mol l−1. By ac impedance spectroscopy the apparent electron transfer rate constant (kapp) of Fe(CN)3−/Fe(CN)4− at the TLA/Au electrode was obtained as 2.5×10−5 cm s−1.  相似文献   

15.
van Staden JF  Stefan RI 《Talanta》1999,49(5):1472-1022
An on-line automated system for the simultaneous flow injection determination of calcium and fluoride in natural and borehole water with conventional calcium-selective and fluoride-selective membrane electrodes as sensors in series is described. Samples (30 μl) are injected into a TISAB II (pH=5.50) carrier solution as an ionic strength adjustment buffer. The sample-buffer zone formed is first channeled to a fluoride-selective membrane electrode and then via the calcium-selective membrane electrode to the reference electrodes. The system is suitable for the simultaneous on-site monitoring of calcium (linear range 10−5–10−2 mol l−1 detection limit 1.94×10−6 mol l−1 recovery 99.22%, RSD<0.5%) and fluoride (linear range 10−5–10−2 mol l−1 detection limit 4.83×10−6 mol l−1 recovery 98.63%, RSD=0.3%) at a sampling rate of 60 samples h−1.  相似文献   

16.
Methanol diffusion in two polymer electrolyte membranes, Nafion 117 and BPSH 40 (a 40% disulfonated wholly aromatic polyarylene ether sulfone), was measured using a modified pulsed field gradient NMR method. This method allowed for the diffusion coefficient of methanol within the membrane to be determined while immersed in a methanol solution of known concentration. A second set of gradient pulses suppressed the signal from the solvent in solution, thus allowing the methanol within the membrane to be monitored unambiguously. Over a methanol concentration range of 0.5–8 M, methanol diffusion coefficients in Nafion 117 were found to increase from 2.9 × 10−6 to 4.0 × 10−6 cm2 s−1. For BPSH 40, the diffusion coefficient dropped significantly over the same concentration range, from 7.7 × 10−6 to 2.5 × 10−6cm2 s−1. The difference in diffusion behavior is largely related to the amount of solvent sorbed by the membranes. Increasing the methanol concentration results in an increase in solvent uptake for Nafion 117, while BPSH 40 actually excludes the solvent at higher concentrations. In contrast, diffusion of methanol measured via permeability measurements (assuming a partition coefficient of 1) was lower (1.3 × 10−6 and 6.4 × 10−7 cm2 s−1 for Nafion 117 and BPSH 40 respectively) and showed no concentration dependence. The differences observed between the two techniques are related to the length scale over which diffusion is monitored and the partition coefficient, or solubility, of methanol in the membranes as a function of concentration. For the permeability measurements, this length is equal to the thickness of the membrane (178 and 132 μm for Nafion 117 and BPSH 40 respectively) whereas the NMR method observes diffusion over a length of approximately 4–8 μm. Regardless of the measurement technique, BPSH 40 is a greater barrier to methanol permeability at high methanol concentrations.  相似文献   

17.
Ohura H  Imato T  Yamasaki S 《Talanta》1999,49(5):1383-1015
A rapid potentiometric flow injection technique for the simultaneous determination of oxychlorine species such as ClO3–ClO2 and ClO3–HClO has been developed, using both a redox electrode detector and a Fe(III)–Fe(II) potential buffer solution containing chloride. The analytical method is based on the detection of a large transient potential change of the redox electrode due to chlorine generated via the reaction of the oxychlorine species with chloride in the potential buffer solution. The sensitivities to HClO and ClO2 obtained by the transient potential change were enhanced 700–800-fold over that using an equilibrium potential. The detection limit of the present method for HClO and ClO2 is as low as 5×10−8 M with use of a 5×10−4 M Fe(III)–1×10−3 M Fe(II) buffer containing 0.3 M KCl and 0.5 M H2SO4. On the other hand, sensitivity to ClO3 was low when a potential buffer solution containing 0.5 M H2SO4 was used, but could be increased largely by increasing the acidity of the potential buffer. The detection limit for ClO3 was 2×10−6 M with the use of a 5×10−4 M Fe(III)–1×10−3 M Fe(II) buffer containing 0.3 M KCl and 9 M H2SO4. By utilizing the difference in reactivity of oxychlorine species with chloride in the potential buffer, a simultaneous determination method for a mixed solution of ClO3–ClO2 or ClO3–HClO was designed to detect, in a timely manner, a transient potential change with the use of two streams of potential buffers which contain different concentrations of sulfuric acid. Analytical concentration ranges of oxychlorine species were 2×10−5–2×10−4 M for ClO3, and 1×10−6–1×10−5 M for HClO and ClO2. The reproducibility of the present method was in the range 1.5–2.3%. The reaction mechanism for the transient potential change used in the present method is also discussed, based on the results of batchwise experiments. The simultaneous determination method was applied to the determination of oxychlorine species in a tap water sample, and was found to provide an analytical result for HClO, which was in good agreement with that obtained by the o-tolidine method and to provide a good recovery for ClO3 added to the sample.  相似文献   

18.
J. Femi Iyun  Ade Adegite 《Polyhedron》1989,8(24):2883-2888
At 25°C, I = 1.0 M (CF3SO3Li++CF3SO3H), [H+] = 0.034–0.274 M and λ = 453 nm, the rate equation for the oxidation of Ti(H2O), 63+ by bromine was found to be: −d/[Br2]T/dt=kK/[Br2][TiIII]/[H+]+K+kK/[Br3][TiIII]/[H++K, where k = 9.2 × 10−3 M −1 s −1 and K = 4.5 × 10−3 M. At [H+] = 1.0 M, [Br] = 0.05–0.4 M, the apparent second-order rate constant decreases as [Br] increases.

The pH-dependence of the oxidation of TiIII-edta by bromine is interpreted in terms of the change in identity of the TiIII-edta species as the pH of the reaction medium changes. The second-order rate constants were fitted using a non-linear least-square computer program with (1/k0edta)2 weighting into an equation of the form: k0edta =k1+k2K1[H+]−1+k3K1K2[H+]−2/1+K1[H+[H+−1+K1K2[H+]−2, with K1 and K2 fixed as earlier determined at 9.55 × 10−3 and 2.29 × 10−9 M, respectively, for the oxidation of bromine. k1=k2=(3.1±0.32)×103M−1s−1 k3=(2.3±0.45)×106N−1s−1.

It is proposed that these electron transfer reactions proceed by univalent changes with the production of Br2.− as a transient intermediate. An outer-sphere mechanism is proposed for these reactions. The homonuclear exchange rate for TiIII-edta+TiIV-edta is estimated at 32 M−1 s−1.  相似文献   


19.
A spectrophotometric study of silver(II) in sulphuric acid solution indicates the formation of two sulphate complexes, in the range 4–18M H2SO4, with absorbance peak maxima at 361 and 260 mμ, respectively. In 15M H2SO4 the molar absorptivity of silver(II) is 3.11 × 104 at 361 mμ. Kinetic studies of the reduction of silver(II) by the solvent suggest a rate-determining step first order in silver(II) and yield a pseudo first-order rate constant of 1.9 × 10−1min−1. Further studies as a function of H2SO4 concentration show that the specific decomposition rate of the two complexes is identical and that changes in H2SO4 concentration only serve to shift the concentration equilibrium between the two complexes.  相似文献   

20.
The second-order rate constants of gas-phase Lu(2D3/2) with O2, N2O and CO2 from 348 to 573 K are reported. In all cases, the reactions are relatively fast with small barriers. The disappearance rates are independent of total pressure indicating bimolecular abstraction processes. The bimolecular rate constants (in molecule−1 cm3 s−1) are described in Arrhenius form by k(O2)=(2.3±0.4)×10−10exp(−3.1±0.7 kJmol−1/RT), k(N2O)=(2.2±0.4)×10−10exp(−7.1±0.8 kJmol−1/RT), k(CO2)=(2.0±0.6)×10−10exp(−7.6±1.3 kJmol−1/RT), where the uncertainties are ±2σ.  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