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.
2.
In situ attenuated total reflectance infrared (ATR-IR) spectroscopy has been applied to the study of the influence of phosphate on the extent of protein adsorption onto TiO2. Immunoglobulin G (Ig.G) was adsorbed onto a TiO2 sol–gel film from solutions containing phosphate or NaCl. Monitoring of the amide II absorbance (v=1545 cm−1) confirmed reduced protein adsorption from the phosphate containing solution. In situ ATR-IR spectroscopy was also used to study phosphate induced desorption of Ig.G. Solutions containing various phosphate concentrations were passed over a TiO2 film with Ig.G adsorbed to it. As the concentration of phosphate increased the amide II absorbance decreased confirming the removal of bound Ig.G from the TiO2 surface. As the amide II absorbance decreased the phosphate absorbance (v=1080 cm−1) increased suggesting accumulation of phosphate at the TiO2 surface. Not all of the bound protein could be displaced from the TiO2 surface by phosphate suggesting the presence of weakly and strongly bound Ig.G.  相似文献   

3.
In this study, phenylalanine as a hydrophobic ligand was covalently attached to the co-monomer methacrylochloride. Then, poly(2-hydroxyethylmethacrylate-co-methacrylamidophenyalanine) [poly(HEMA–MAPA)] membranes were prepared by UV-initiated photopolymerization of HEMA and methacrylamidophenyalanine. The γ-globulins adsorption onto these affinity membranes from aqueous solutions containing different amounts of γ-globulins at different pH was investigated in a batch system. The γ-globulins adsorption capacity of the membranes was increased as the ligand density on the membrane surface increase. The non-specific adsorption of the γ-globulins on the pHEMA membranes was negligible. The adsorption phenomena appeared to follow a typical Langmuir isotherm. The maximum adsorption capacity (qm) of the poly(HEMA–MAPA4) membrane for γ-globulins was 2.37 mg g−1 dry membrane. The equilibrium constant (kd) value was found to be 1.61×10−1 mg ml−1. More than 87% (up to 100%) of the adsorbed γ-globulins were desorbed in 120 min in the desorption medium containing 50% ethylene glycol in 1.0 M NaCl.  相似文献   

4.
Bilayer lipid membranes (BLM) are commonly used as models for cell membranes to study their interactions with inorganic ions and molecules of biological importance. In this work the principal electrostatic effects at the BLM surface are demonstrated by two methods: by the inner membrane field compensation (IFC) which is applied to planar BLM and sensitive to changes in the total boundary potential; and by electrokinetic measurements in liposome suspensions, sensitive to diffuse (surface) component of this potential. The difference in these two potentials allows us to conclude on changes in the dipole component of the boundary potential caused by structural changes at the membrane–water interface. No difference in the experimental data of both methods was observed for Be2+ and other divalent cation adsorption to unchanged phosphatidyl choline (PC) membranes. These data are in a good agreement with the Gouy–Champan–Stern (GCS) theory of diffuse double layer. This theory gives the value of binding constants for Be2+ about 400 M−1 and 104 M−1 for DPPC liposomes in the liquid and solid states of the lipids, respectively. Clear isotope effects for normal and heavy water solutions of Be2+ were observed both by the electrostatic measurements and by differential scanning calorimetry. In contrast to PC, the electrostatic potentials induced by Be2+ and Gd3+ adsorption to membranes from phosphatidyiserine (PS) show the difference between the data of mentioned methods — total boundary potential changes are much higher in comparison to the surface potential. Dipole potential changes (about 150 mV) caused by changes in PS head group orientation may be more important in this specific case.  相似文献   

5.
The surface tension isotherms for pure oligooxypropylenated piperidine and morpholine at the aqueous solution—air interface were determined and interpreted. The surface excess concentration, Γ, the surface area per molecule, A, and the standard free energy of adsorption, ΔG°, were calculated according to a new empirical adsorption equation. The standard free energy contribution for the oxypropylene group (PO) in morpholine derivatives,ΔG° (PO) = −3.34 kJ mol−1, is substantially lower than that for the PO group located in the piperidine derivatives, i.e. ΔG° (PO)= −3.12 kJ mol−1.  相似文献   

6.
The effects of external stimuli such as pH of the buffer solution, ionic strength, temperature and the amount of poly-electrolyte monomer in the hydrogel system on the Bovine Serum Albumin (BSA) adsorption capacity of poly(acrylamide/maleic acid) [P(AAm/MA)] hydrogels were investigated. Poly-electrolyte P(AAm/MA) hydrogels with varying compositions were prepared by irradiating acrylamide/maleic acid/water mixtures with γ rays at ambient temperature. Langmuir type adsorption isotherms were observed for all prepared hydrogels. Increase of ionic strength of the buffer solution from 0.01 to 0.1 mol dm−3 decreased the adsorption capacity of hydrogels and zero adsorption was observed in the presence of 0.1 mol dm−3 Na+ and Ca2+ ion in the adsorption medium. The adsorption capacity of hydrogels was found to increase from 0 to 120 mg BSA/g dry gel, by changing external stimuli and hydrogel composition.  相似文献   

7.
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   

8.
Dipalmitoylphosphatidylcholine (DPPC) monolayers were deposited onto a germanium attenuated total reflectance (ATR) crystal using the Langmuir–Blodgett technique. The DPPC-coated crystal was then exposed to human serum albumin or human fibrinogen solutions while measuring the protein adsorption by recording FTIR spectra. The effect of the zwitterionic nature of the DPPC polar headgroup towards protein adsorption has been ascertained by exposing either the phospholipid headgroup or the acyl chains to the protein solution; this was possible by the use of a silanized or a bare germanium crystal. Calibration curves have been made to measure the protein surface concentrations. After 3 h, the albumin surface concentration on DPPC monolayers was about three times higher when the proteins were exposed to the lipid acyl chains instead of the polar headgroups (e.g. 3 vs. 1 μg cm−2). As for fibrinogen (FGN) adsorption, when the lipid polar headgroups were exposed to the protein solution, the FGN adsorption was low reaching a maximum value of 0.5 μg cm−2. When interacting with the lipid acyl chains, the FGN adsorption reached a plateau at a value of 2.1 μg cm−2 after 3 h. Clearly, both albumin and FGN showed a low tendency to adsorb on surfaces where the lipid polar headgroups are exposed toward the protein solution.  相似文献   

9.
A flow injection on-line sorption preconcentration electrothermal atomic absorption spectrometric system for fully automatic determination of lead in water was investigated. The discrete non-flow-through nature of ETAAS, the limited capacity of the graphite tube and the relatively large volume of the knotted reactor (KR) are obstacles to overcome for the on-line coupling of the KR sorption preconcentration system with ETAAS. A new FI manifold has been developed with the aim of reducing the eluate volume and minimizing dispersion. The lead diethyldithiocarbamate complex was adsorbed on the inner walls of a knotted reactor made of PTFE tubing (100 cm long, 0.5 mm i.d.). After that, an air flow was introduced to remove the residual solution from the KR and the eluate delivery tube, then the adsorbed analyte chelate was quantitatively eluted into a delivery tube with 50 μl of ethanol. An air flow was used to propel the eluent from the eluent loop through the reactor and to introduce all the ethanolic eluate onto the platform of the transversely heated graphite tube atomizer, which was preheated to 80°C. With the use of the new FI manifold, the consumption of eluent was greatly reduced and dispersion was minimized. The adsorption efficiency was 58%, and the enhancement factor was 142 in the concentration range 0.01–0.05 μg l−1 Pb at a sample loading rate of 6.8 ml min−1 with 60 s preconcentration time. For the range 0.1–2.0 μg l−1 of Pb a loading rate of 3.0 ml min−1 and 30 s preconcentration time were chosen, resulting in an adsorption efficiency of 42% and an enhancement factor of 21, respectively. A detection limit (3σ) of 2.2 ng l−1 of lead was obtained using a sample loading rate of 6.8 ml min−1 and 60 s preconcentration. The relative standard deviation of the entire procedure was 4.9% at the 0.01 μg l−1 Pb level with a loading rate of 6.8 ml min−1 and 60 s preconcentration, and 2.9% at the 0.5 μg l−1 Pb level with a 3.0 ml min−1 loading rate and 30 s preconcentration. Efficient washing of the matrix from the reactor was critical, requiring the use of the standard addition method for seawater samples. The analytical results obtained for seawater and river water standard reference materials were in good agreement with the certified values.  相似文献   

10.
Zenki M  Tanishita A  Yokoyama T 《Talanta》2004,64(5):1273-1277
Ascorbic acid (AA) could be determined in large quantities of a co-existing oxidant. The incorporation of an on-line reagent regeneration step based on redox reaction eliminates the baseline drift in the procedure. This makes it possible to adopt a circulatory flow injection method (cyclic FIA) and to determine AA repetitively. The method is based on the reduction of iron(III) to iron(II) by the analyte, the reaction of the produced iron(II) with 1,10-phenanthroline (phen) in a weak acidic medium to form a colored complex, and the subsequent oxidation reaction of iron(II) to iron(III) by the co-existing peroxodisulfate. A solution (50 ml) of 3.0×10−4 mol l−1 ferric chloride, 9.0×10−4 mol l−1 phen and 5.0×10−2 mol l−1 ammonium peroxodisulfate in acetate buffer (0.2 mol l−1, pH 4.5) is continuously circulated at a constant flow rate of 1.0 ml min−1. Into this stream, an aliquot (20 μl) of the sample solution containing AA is quickly injected by means of a six-way valve. The complex formed is monitored spectrophotometrically (at 510 nm) in the flow system. The stream then returns to the reservoir after passing through a time-delay coil (50 m). The iron(II)–(phen)3 complex is oxidized to iron(III)–(phen)3 complex by peroxodisulfate which exists excessively in the circulating reagent solution. The proposed method allows as many as 300 repetitive determinations of 15 mg l−1 AA with only 50 ml reservoir solution. The contents of AA in commercial pharmaceutical products were analyzed to demonstrate the capability of the developed system.  相似文献   

11.
Electrophoretic mobility and contact angle measurements have been made on alatrofloxacin mesylate and its formulations which were protected from or exposed to light, and its degradation product compound (F). In aqueous solution, the light-protected alatrofloxacin mesylate had a zeta-potential of +19 mV, a negligible electron-acceptor (γi+) surface tension parameter and an electron-donor surface tension parameter γi=32.5 mJ m−2, which was higher than that of water. This caused the particles to be very hydrophilic and to form very stable suspensions in aqueous solution due, mainly, to a net Lewis acid–base (polar) repulsion. After the suspensions were exposed to light, the zeta-potential of the degradation product increassed to +37.8 mV, but the electron-donor surface tension parameter decreased to γi=8 mJ m−2, making the molecules or particles very hydrophobic and causing them to flocculate. The energies of attraction in the latter case were mainly hydrophobic (90%) with about 10% resulting from van der Waals forces.  相似文献   

12.
The paper reports results of a study on the specific adsorption of F, Cl, Br, I, ClO3, BrO3, IO3 and IO4 on hydrous γ-Al2O3. The isotherms of the anion adsorption and the adsorption dependencies on pH and the ionic strength of the solution have been determined under the equilibrium conditions. According to the degree of affinity to γ-Al2O3, the anions can be ordered as: I3334−. It has been established that the sorption of IO4 and F involves the formation of surface complexes in the inner co-ordination sphere, whereas that of Cl, Br, I, ClO3, BrO3 and IO3 takes place through formation of ion pair complexes in the outer co-ordination sphere. In the dynamic system, the exchange isoplanes and elution curves have been determined for selected anions on columns filled with Al2O3. It has been shown that γ-Al2O3 can be used for isolation and concentration of IO3 from natural waters in order to decrease the limit of the ions determination to 2 μg l−1. Using differential pulse voltammetry (DPV), after isolation and concentration on γ-Al2O3, the content of iodates has been determined in mineral, marine and tap water doped with these ions.  相似文献   

13.
A sample solution was passed at 20 ml min−1 through a column (150×4 mm2) of Amberlite IRA-410Stron anion-exchange resin for 60 s. After washing, a solution of 0.1% sodium borohydride was passed through the column for 60 s at 5.1 ml min−1. Following a second wash, a solution of 8 mol l−1 hydrochloric acid was passed at 5.1 ml min−1 for 45 s. The hydrogen selenide was stripped from the eluent solution by the addition of an argon flow at 150 ml min−1 and the bulk phases were separated by a glass gas–liquid separator containing glass beads. The gas stream was dried by passing through a Nafion® dryer and fed, via a quartz capillary tube, into the dosing hole of a transversely heated graphite cuvette containing an integrated L’vov platform which had been pretreated with 120 μg of iridium as trapping agent. The furnace was held at a temperature of 250°C during this trapping stage and then stepped to 2000°C for atomization. The calibration was performed with aqueous standards solution of selenium (selenite, SeO32−) with quantification by peak area. A number of experimental parameters, including reagent flow rates and composition., nature of the gas–liquid separator, nature of the anion-exchange resin, column dimensions, argon flow rate and sample pH, were optimized. The effects of a number of possible interferents, both anionic and cationic were studies for a solution of 500 ng 1−1 of selenium. The most severe depressions were caused by iron (III) and mercury (II) for which concentrations of 20 and 10 mg  1−1 caused a 5% depression on the selenium signal. For the other cations (cadmium, cobalt, copper, lead,. magnesium, and nickel) concentrations of 50–70 mg 1−1 could be tolerated. Arsenate interfered at a concentration of 3 mg−1, whereas concentrations of chloride, bromide, iodide, perchlorate, and sulfate of 500–900 mg l−1 could be tolerated. A linear response was obtained between the detection limit of 4 ng 1−1, with a characteristic mass of 130 pg. The RSDs for solutions containing 100 and 200 ng 1−1 selenium were 2.3% and 1.5%, respectively.  相似文献   

14.
The cathodic adsorptive electrochemical behavior of guanine in the presence of some metal ions at the static mercury drop electrode was investigated. A 1.0×10−3 mol l−1 NaOH or a 2.0×10−2 mol l−1 Hepes buffer at pH 8.0 solutions were used as supporting electrolytes. The reduction peak potential for guanine was found to be around −0.15 V, which is very close to the mercury reduction wave. A new peak appears at −0.60 V in the presence of copper or at −1.05 V in the presence of zinc. A square wave voltammetric procedure for electroanalytical determination of guanine in 2.0×10−2 mol l−1 Hepes buffer at pH 8.0 containing 1.6×10−5 mol l−1of copper ions, was developed. An accumulation potential of −0.15 V during 270 s for the prior adsorption of guanine at the electrode surface was used. The response of the system was found to be linear in the range of guanine concentration from 6.62×10−8 to 1.32×10−7 mol l−1 and the detection limit was 7.0×10−9 mol l−1. The influence of DNA bases such as adenine, cytosine and thymine was also examined. Cyclic voltammetry was used to characterize the interfacial and redox mechanism.  相似文献   

15.
Thermodynamic characterisation of the adsorption process (at a low temperature) of dihydrogen on the zeolite Li-ZSM-5 was carried out by means of variable-temperature infrared spectroscopy, with the simultaneous measurement of temperature and equilibrium pressure. Adsorption renders the H–H stretching mode infrared active, at 4092 cm−1. The standard adsorption enthalpy and entropy resulted to be ΔH0=−6.5(±0.5) kJmol−1 and ΔS0=−90(±5) Jmol−1 K−1, respectively. The adsorption enthalpy is significantly larger than the liquefaction heat, and this fact renders Li-ZSM-5 a potential cryoadsorbent for hydrogen storage.  相似文献   

16.
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   

17.
The rate constants at which oxidizing and reducing radicals react with the dinuclear iron(III) complex Fe2O(ttha)2− were measured in neutral aqueous solution. The rate constants for reduction of the complex by ·CO2.− CH3.CHOH and O2.− were found to be comparable with rate constants previously measured in mononuclear iron(III) polyaminocarboxylate systems. Fe2O(ttha)2− reacts slowly with O2.− (k8 = (1.2 ± 0.2) × 104 dm3 mol−1 s−1) and, hence, is a relatively poor catalyst for the dismutation of superoxide radical. The hydrated electron reduces the complex at a diffusion-controlled rate in a process which consumes one proton: eaq + Fe2O(ttha)2− → Fe2III,IIO(ttha)3− The reduction by carbon-centered radicals produces a (III,II) mixed-valence complex with an absorption spectrum different from that of the Fe2(II,III) species produced from reduction by the hydrated electron. The oxidizing radicals .OH and ·CO3 appear to act as reductants of the complex via ligand oxidation rather than by oxidation of the Fe2IIIO core to Fe2III,IVO. In the former case ligand attack appears to occur mainly at the methylene carbon of a glycinate group. The decarboxylation product, CO2, was detected by its aquation reaction in the presence of a pH sensitive dye, bromthymol blue.  相似文献   

18.
A reversed flow injection colorimetric procedure for determining iron(III) at the μg level was proposed. It is based on the reaction between iron(III) with norfloxacin (NRF) in 0.07 mol l−1 ammonium sulfate solution, resulting in an intense yellow complex with a suitable absorption at 435 nm. Optimum conditions for determining iron(III) were investigated by univariate method. The method involved injection of a 150 μl of 0.04% w/v colorimetric reagent solution into a merged streams of sample and/or standard solution containing iron(III) and 0.07 mol l−1 ammonium sulfate in sulfuric acid (pH 3.5) solution which was then passed through a single bead string reactor. Subsequently the absorbance as peak height was monitored at 435 nm. Beer's law obeyed over the range of 0.2–1.4 μg ml−1 iron(III). The method has been applied to the determination of total iron in water samples digested with HNO3–H2O2 (1:9 v/v). Detection limit (3σ) was 0.01 μg ml−1 the sample through of 86 h−1 and the coefficient of variation of 1.77% (n=12) for 1 μg ml−1 Fe(III) were achieved with the recovery of the spiked Fe(III) of 92.6–99.8%.  相似文献   

19.
Rhodium particles in nanometer size were prepared by impregnating alumina powders with aqueous solutions containing rhodium salts. The dispersion (D) of rhodium crystallites on the prepared samples was estimated by dioxygen adsorption measured at 300 K. Phenomena of oxidizing the supported crystallites with 2.5 × 104 Pa O2 in a temperature range between 280 and 870 K were calorimetrically studied. Extent of oxidation may be distinguished into three stages, i.e., adsorption on surface (T < 300 K), progressive penetration into bulk, and formation of a stable bulk oxide (T> 700 K), on raising the oxidation temperature. Heat of dioxygen adsorption varies only slightly with the dispersion (D) of rhodium and has a value of 294 ± 6 kJ (mol O2)−1. Chemical stoichiometry of the bulk oxide formed, however, varies with the dispersion of rhodium crystallites. A dioxide (RhO2) (f H = 225 ± 3 kJ (mol O2)−1) and a sesquioxide (Rh2O3) (f H = 273 ± 3 kJ (mol O2)−1) was formed at D < 60% and D> 80%,  相似文献   

20.
Liposomes can be effectively deposited on the inner surface of a capillary wall by flushing the electrophoretic system with a liposome suspension followed by air-drying of the capillary and removal of the excess of loosely bound liposomes by a 0.1 M NaOH wash. It was demonstrated that capillaries prepared in this way could be used for studies of analyte (drug)–liposome binding. The results were expressed as free binding energy changes [Δ(ΔG0)] relatively to an arbitrarily selected standard (acetylsalicylic acid). The results were compared to [Δ(ΔG0)] changes obtained from binding studies effected by capillary electrophoresis using a stable liposome plug in a capillary with minimized endoosmotic flow. Good agreement of data reported in the literature (without correction for the residual endoosmotic flow), our previous data obtained in a similar way (however, after the correction for the residual endoosmotic flow) and data obtained by the immobilized liposome affinity electrochromatography reported in this communication was achieved.  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