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.
Ultrathin nylon capsule membranes coated with synthetic bilayers, the hydrophilic head groups of which had cationic, anionic, zwitterionic and nonionic charges, were prepared. Permeation of NaCl trapped in the inner aqueous phase was reduced by a factor of 10-1000 relative to that of the uncoated, semipermeable capsules, and drastically changed at the phase transition temperature, Tc, of the coating bilayers, depending on the charge of their hydrophilic head group. In the case of capsules coated with positively or negatively charged bilayers, NaCl permeation was enhanced at temperatures above the Tc of the coating bilayers, as expected. On the other hand, NaCl release of capsules coated with neutral charged (nonionic and zwitterionic) bilayers was largely reduced at temperatures above the Tc. From activation energy data of Arrhenius plots, the permeation mechanism of NaCl, depending on the membrane surface charge, below and above the Tc was discussed.  相似文献   

2.
The hydrolytic activity of phospholipase A2 (PLA2) against the dipalmitoylphosphatidylcholine monolayer formed at the nitrobenzene-water interface has been studied under the control of the potential drop across the monolayer. The activities of both porcine pancreatic and Naja naja PLA2S was the highest when the potential of the nitrobenzene phase was 60 mV negative with respect to that of the aqueous phase. The local electrostatic interaction between the positively charged domain, the recognition site, of PLA2 molecules with the negatively charged substrate side of the interface, where zwitterionic substrate molecules and negatively charged product molecules were adsorbed, is an important factor in determining the interfacial enzymatic activity. Irreversible adsorption of PLA2 molecules on the substrate monolayer is confirmed, giving unequivocal evidence for the scooting mode of hydrolysis by PLA2.  相似文献   

3.
The ternary phase diagrams of zwitterionic single-chain hydrocarbon surfactant (tetradecyldimethylaminoxide, C14DMAO)—a perfluoro cosurfactant (1,1-H-dihydroperfluorooctanol, C7F15CH2OH)-H2O, and C14DMAO-C7F15CH2OH-H2O-HCl have been studied at 25°C. The identification of the phases was done by means of electrical conductivity, optical polarizing microscopy, and 2H-NMR techniques. In this system the originally uncharged zwitterionic surfactant was increasingly charged by protonation through addition of HCl. The sequence of the phases observed is similar to that observed for hydrocarbon surfactant–cosurfactant mixtures, namely, L1, L1/Lα, Lα and L2, as relative volume fraction of apolar compound increases over that of polar compound. The absence of sponge phase (L3) is a result of the high bending constant of the mixed bilayers in this perfluoro system.  相似文献   

4.
Three unimolecular peptide channels have been designed and prepared by using the β‐helical conformation of gramicidin A (gA). The new peptides bear one to three NH3+ groups at the N‐end and one to three CO2? groups at the C‐end. These zwitterionic peptides were inserted into lipid bilayers in an orientation‐selective manner. Conductance experiments on planar lipid bilayers showed that this orientation bias could lead to observable directional K+ transport under multi‐channel conditions. This directional transport behavior can further cause the generation of a current across a planar bilayer without applying a voltage. More importantly, in vesicles with identical external and internal KCl concentrations, the channels can pump K+ across the lipid bilayer and cause a membrane potential.  相似文献   

5.
It is widely known that ions in the aqueous phase affect the binding of charged solutes to membranes. Here we report the effect of ions and urea on the interaction of both the charged and uncharged forms of the local anesthetic tetracaine (TTC, an aminoester derivative of ben-zoic acid) to zwitterionic micelles. Binding was monitored by the increase in TTC fluorescence. Shifts in the emission wavelength maximum (δmax) indicated that the anesthetic was located in an environment of lower polarity. The neutral form of TTC bound to micelles to a larger extent than the protonated form, in agreement with results found for lipid bilayers (Boulanger, Leitch, Schreier and Smith, Can. J. Biochem. 58 , 986–995, 1980). When ions from the Hofmeister series and urea were compared in their ability to affect the partitioning of the anesthetic, binding of both the charged and uncharged forms was found to increase upon addition of SO42– and CI? but was seen to decrease in the presence of SCN?, CIO4?- and urea. Solubility measurements revealed that the solubility of uncharged TTC increases in solutions containing the additives in the following order: SO42– < CI? < CIO4?- < dilute buffer < SCN? < urea. Spin label EPR spectra indicated that, except for CIO4?-, the ions had little effect on micellar structure. Static light scattering measurements corroborated this result indicating a large increase in micellar molecular weight in the presence of CIO4?- and lesser increases for CI.- and SCN?. The results show that, besides affecting the binding of ionic species through an electrostatic mechanism, ions also act by altering water structure and, as a consequence, the water solubility and the tendency to partition into the less polar micellar environment of polar charged or uncharged small organic solutes such as the benzoic acid ester derivative. Moreover, evidence suggests that the ions bind directly to the zwitterionic polar groups of the micelles, leading to changes in structure and size.  相似文献   

6.
Bilayer swelling behavior of nonionic and anionic surfactant mixed aqueous solution induced by adding glycerin was studied. The phenomenon were performed on a system, polyethylene glycol ether of tridecyl alcohol with the average number of ethylene oxide of 5 (CH3(CH2)12(OCH2CH2)5OH; abbreviation IT5) and SDS mixed aqueous solution, with white cream of the upper phase and micelles (L1) of the lower phase. White cream containing densely packed multilamellar vesicles was revealed by freeze-fracture transmission electron microscopy and polarized microscope observations. Phase transition from white cream/L1, two-phase, to clear unique vesicle phase can be induced by adding glycerin to replace water. The addition of glycerin lowers the turbidity of the dispersion and swells the interlamellar distance between bilayers, which could be explained by refractive-index matching between solvent and bilayers.  相似文献   

7.
In the title 2:1 salt, 2C2H6NO2+·C2O42−, the glycine mol­ecule is in the cationic form with a positively charged amino group and an uncharged carboxylic acid group. The doubly charged oxalate anion lies across a crystallographic inversion centre. One of the reasons why the 1:1 glycinium oxalate salt has a higher melting point than the title compound may be the difference in their hydrogen‐bonding patterns. A database search for salts formed between amino acids or substituted amino acids and oxalic acid revealed that, in most of the structures, the conformation about the O=C—OH bond is synplanar. d ‐Tryptophan oxalate is the only example where the OH group of a semi‐oxalate adopts an anti­planar conformation. The 2:1 stoichiometry seen in the present salt is observed only in the salts of dl ‐serine, dl ‐aspartic acid and betaine with oxalic acid.  相似文献   

8.
The anionization of molecules supplied from the gas phase onto a negatively charged [Cl]? or [NO3]? ion donating surface has been investigated. The charged surface was prepared by exposing an aqueous solution of LiCl (or LiNO3) and polyethylene oxide to a high external field as is done in negative ion field desorption mass spectrometry. The ionization of some monosaccharides and adenosine by [Cl]? and [NO3]? attachment and of some acids by proton abstraction is reported.  相似文献   

9.
The radiolysis of aqueous vitamin B5 (pantothenic acid) has been investigated under various experimental conditions. The highest vitamin degradation (G=3.22) was observed in solutions saturated with N2O, where 90% OH radicals are operating. As final products, the following were established: aldehydes, carboxylic acids and ammonia. Their yield strongly depends on the presence/absence of air as well as on N2O (used to convert eaq into OH) and was determined as a function of absorbed radiation dose. HPLC-analysis showed that in all media, a main product is formed, having the highest yield in aerated solutions.Based on the chemical analysis, it appears that the OH radicals are most involved in the degradation process. A precise sequence of the reaction steps could not be given presently, because of the implication of many simultaneous reactions.  相似文献   

10.
The apparent molar volume, V o φ, 2, of glycine, alanine, α-amino-n-butyric acid, valine and leucine have been determined in aqueous solutions of 0.25, 0.5 and 1.0 mol⋅dm−3 magnesium sulfate, and the partial specific volume from density measurements at 298.15 K. These data have been used to calculate the infinite dilution apparent molar volume, V o 2,m , group contribution of amino acids and partial molar volume of transfer, Δtr V 2,m o, from water to aqueous magnesium sulfate solutions. The linear correlation of V 2,m o for a homologous series of amino acids has been utilized to calculate the contributions of charged end groups (NH3 +, COO), CH2 - groups and other alkyl chains of amino acids to V 2,m o. The results for Δtr V 2,m o of amino acids from water to aqueous magnesium sulfate solutions have been interpreted in terms of ion-ion, ion-polar, hydrophilic-hydrophilic and hydrophobic-hydrophobic group interactions. The values of the standard partial molar volume of transfer for the amino acids with different hydrophobic contents, from water to aqueous MgSO4 are in general positive, indicating the predominance of the interactions of zwitterionic/hydrophilic groups of amino acids with ions of the salt. The hydration number decreases with increasing concentration of salt. The number of water molecules hydrated to amino acids decreases, further strengthening the predominance of ionic/hydrophilic interactions in this system.  相似文献   

11.
The principle of novel flow-through sensor systems with FTIR spectroscopic detection is presented on the example of the determination of organic acids in aqueous solution. The constructed flow-through sensor system is based on trapping of derivatized porous polymer beads in a conventional IR transmission cell and integration of the flow cell into a sequential injection (SI) manifold. By the SI-manifold sample pre-conditioning, sample-sensor interaction and sensor regeneration were performed in an automated and highly reproducible way. The polymer beads used in this study contained anion exchanger groups so that negatively charged molecules such as organic acids present in the anionic form could selectively interact with the polymer beads. Upon pumping a sample through the sensor cell organic acids were retained on the polymer beads whereas non-ionic matrix molecules passed hence allowing to separate the target analytes form the matrix. Apart from that the organic acids were also concentrated onto the polymer beads so that absolute analyte amounts in the low μg range could easily be detected. Linear calibration curves from 0 to 1 mmol l−1 were recorded for acetic and malic acid using a sample volume of 500 μl (sx0: 0.032 mmol l−1 acetic acid and 0.031 mmol l−1 malic acid). Mixtures of both acids were analyzed as well and it could be shown that by application of multivariate data evaluation procedures (PLS) simultaneous quantification of both acids could be performed successfully using the developed flow-through sensor system.  相似文献   

12.
Paramagnetic products of low-temperature X-ray radiolysis of aqueous poly(vinyl alcohol) solutions (2.5 and 5% by weight) were studied by ESR spectroscopy. Experimental spectra were ascribed to a superposition of signals from hydroxyl radicals and –CH2??C(OH)–CH2? macroradicals (Cα-macroradicals), respectively. No ESR signals corresponding to trapped electrons were observed that was attributed to the peculiarities of microheterogenous structure of the frozen aqueous polymer solutions. Annealing at 115 K resulted in partial conversion of OH radicals to Cα-macroradicals. It was suggested that main part of hydroxyl radicals was stabilized in phase of polycrystalline ice while macroradicals were formed in “mixed” water–polymer phase. The radiation–chemical yields of paramagnetic species stabilized in the systems under study were determined.  相似文献   

13.
The apparent molar volumes Vφ of glycine, alanine, valine, leucine, and lysine have been determined in aqueous solutions of 0.05, 0.5, 1.0 mol · kg−1 sodium dodecyl sulfate (SDS) and 1.0 mol · kg−1 cetyltrimethylammonium bromide (CTAB) by density measurements at T=298.15 K. The apparent molar volumes have also been determined for diglycine and triglycine in 1 mol · kg−1 SDS and CTAB solutions. These data have been used to calculate the infinite dilution apparent molar volumes V20 for the amino acids and peptides in aqueous SDS and CTAB and the standard partial molar volumes of transfer (ΔtrV2,m0) of the amino acids and peptides to these aqueous surfactant solutions. The linear correlation of V20 for a homologous series of amino acids has been utilized to calculate the contribution of the charged end groups (NH3+, COO), CH2 group and other alkyl chains of the amino acids to V20. The results on the partial molar volumes of transfer from water to aqueous SDS and CTAB have been interpreted in terms of ion–ion, ion–polar and hydrophobic–hydrophobic group interactions. The volume of transfer data suggests that ion–ion or ion–hydrophilic group interactions of the amino acids and peptides are stronger with SDS compared to those with CTAB. Comparison of the hydration numbers of amino acids calculated in the present studies with those in other solvents from literature shows that these numbers are almost the same at 1 mol · kg−1 level of the cosolvent/cosolute. Increasing molality of the cosolvent/cosolute beyond 1 mol · kg−1 lowers the hydration number of the amino acids due to increased interactions with the solvent and reduced electrostriction.  相似文献   

14.
15.
We report the formation and characterization of multiply negatively charged sodium bis(2-ethylhexyl) sulfosuccinate (NaAOT) aggregates in the gas phase, by electrospray ionization of methanol/water solution of NaAOT followed by detection using a guided-ion-beam tandem mass spectrometer. Singly and doubly charged aggregates dominate the mass spectra with the compositions of [Nan-zAOTn]z– (n?=?1–18 and z?=?1–2). Solvation by water was detected only for small aggregates [Nan-1AOTnH2O] of n?=?3–9. Incorporation of glycine and tryptophan into [Nan-zAOTn]z– aggregates was achieved, aimed at identifying effects of guest molecule hydrophobicity on micellar solubilization. Only one glycine molecule could be incorporated into each [Nan-zAOTn]z– of n?≥?7, and at most two glycine molecules could be hosted in that of n?≥?13. In contrast to glycine, up to four tryptophan molecules could be accommodated within single aggregates of n?≥?6. However, deprotonation of tryptophan significantly decrease its affinity towards aggregates. Collision-induced dissociation (CID) was carried out for mass-selected aggregate ions, including measurements of product ion mass spectra for both empty and amino acid-containing aggregates. CID results provide a probe for aggregate structures, surfactant-solute interactions, and incorporation sites of amino acids. The present data was compared with mass spectrometry results of positively charged [Nan+zAOTn]z+ aggregates. Contrary to their positive analogues, which form reverse micelles, negatively charged aggregates may adopt a direct micelle-like structure with AOT polar heads exposed and amino acids being adsorbed near the micellar outer surface.   相似文献   

16.
《Fluid Phase Equilibria》1999,155(2):251-259
Osmotic coefficient data were obtained for the aqueous solutions of NaOH–NaCl–NaAl(OH)4. The solutions were prepared by dissolving AlCl3·6H2O in aqueous NaOH solutions. The osmotic coefficients of the solutions were measured by an isopiestic method at 25°C. The osmotic coefficient data were used to evaluate the unknown binary and mixing parameters of Pitzer's model for the aqueous NaOH–NaCl–NaAl(OH)4–H2O system. The binary Pitzer's parameters, β(0), β(1), and Cφ, for NaAl(OH)4 were found to be −0.0083, 0.0710, and 0.00184 respectively. These binary parameters were obtained from the data on the ternary system. This was necessary since it was not possible to prepare a single (NaAl(OH)4) solution. The mixing parameters, ΘOHAl(OH)4, ΘClAl(OH)4, ΨNa+OHAl(OH)4, and ΨNa+ClAl(OH)4 were found to be −0.2255, −0.2430, −0.0388, and 0.2377 respectively. The experimental osmotic coefficient data were correlated well with Pitzer's model using the parameters obtained.  相似文献   

17.
Isotherms of copper cation sorption by H-ZSM-5 zeolite from aqueous and aqueous ammonia solutions of copper acetate, chloride, nitrate, and sulfate are considered in terms of Langmuir’s monomolecular adsorption model. Using UV-Vis diffuse reflectance spectroscopy, IR spectroscopy, and temperatureprogrammed reduction with hydrogen and carbon monoxide, it has been demonstrated that the electronic state of the copper ions is determined by the ion exchange and heat treatment conditions. The state of the copper ions has an effect on the redox properties and reactivity of the Cu-ZSM-5 catalysts in the selective catalytic reduction (SCR) of NO with propane and in N2O decomposition. The amount of Cu2+ that is sorbed by zeolite H-ZSM-5 from aqueous solution and is stabilized as isolated Cu2+ cations in cationexchange sites of the zeolite depends largely on the copper salt anion. The quantity of Cu(II) cations sorbed from aqueous solutions of copper salts of strong acids is smaller than the quantity of the same cations sorbed from the copper acetate solution. When copper chloride or sulfate is used, the zeolite is modified by the chloride or sulfate anion. Because of the presence of these anions, the redox properties and nitrogen oxides removal (DeNO x ) efficiency of the Cu-ZSM-5 catalysts prepared using the copper salts of strong acids are worse than the same characteristics of the sample prepared using the copper acetate solution. The addition of ammonia to the aqueous solutions of copper salts diminishes the copper salt anion effect on the amount of Cu(II) sorbed from these solutions and hampers the nonspecific sorption of anions on the zeolite surface. As a consequence, the redox and DeNO x properties of Cu-ZSM-5 depend considerably on the NH4OH/Cu2+ ratio in the solution used in ion exchange. The aqueous ammonia solutions of the copper salts with NH4OH/Cu2+ = 6–10 stabilize, in the Cu-ZSM-5 structure, Cu2+ ions bonded with extraframework oxygen, which are more active in DeNO x than isolated Cu2+ ions (which form at NH4OH/Cu2+ = 30) or nanosized CuO particles (which form at NH4OH/Cu2+ = 3).  相似文献   

18.
A new ion chromatographic (IC) system, in which n-tetradecylphosphocholine (TDPC, a phosphobetaine type of zwitterionic surfactant) was used as the stationary phase, pure water as the mobile phase, and conductivity as the method of detection, has been developed for the determination of inorganic acids. Five model acids, HCl, HNO3, HClO4, H2SO4, and H3PO4, were separated to baseline and eluted in the order H3PO4 > HCl > HNO3 > H2SO4 > HClO4. When peak areas were plotted against the concentrations of the acids in samples, linear calibration curves were obtained. Ultimate determination limits were approximately 1 mmol L–1, but the discrimination of the method between solutions of different concentration was better than 10 μmol L–1 for those model analytes. Salts of divalent cations could also be separated, but they were eluted faster than the acids. No separation was observed for the salts of monovalent cations. This newly proposed approach is applicable to the simultaneous determination of the inorganic acids (produced by reactions of NOx, SOx, and HCl with water) in aerosols.  相似文献   

19.
The acid–base chemistry of three zwitterionic compounds, namely anthranilic (2-aminobenzoic acid), N-methylanthranilic and N-phenylanthranilic acid has been characterized in terms of the macroconstants K a1, K a2, the isoelectric point pH I, the tautomerization constant K z and microconstants k 11, k 12, k 21, k 22. The potentiometric titration method was used to determine the macrodissociation constants. Due to the very poor water solubility of N-phenylanthranilic acid the dissociation constants pK a1 and pK a2 were determined in MDM–water mixtures [MDM is a co-solvent mixture, consisting of equal volumes of methanol (MeOH), dioxane and acetonitrile (MeCN)]. The Yasuda–Shedlovsky extrapolation procedure has been used to obtain the values of pK a1 and pK a2 in aqueous solutions. The pK a1 and pK a2 values obtained by this method are 2.86 ± 0.01 and 4.69 ± 0.03, respectively. The tautomerization constant K z describing the equilibrium between unionized form ? zwitterionic form was evaluated by the K z method based on UV–VIS spectrometry. The method uses spectral differences between the zwitterionic form (found at isoelectric pH in aqueous solution) and the unionized form (formed in an organic solvent of low dielectric constant). The highest value of the K z constant has been observed in the case of N-methylantranilic acid (log10 K z = 1.31 ± 0.04). The values of log10 K z for anthranilic and N-phenylanthranilic acids are similar and have values of 0.93 ± 0.03 and 0.90 ± 0.05, respectively. The results indicate that the tested compounds, in aqueous solution around the isoelectric point pHI, occur mainly in the zwitterionic form. Moreover, the influence of the type of substituent and pH of the aqueous phase on the equilibrium were analyzed with regard to the formation and the coexistence of different forms of the acids in the examined systems.  相似文献   

20.
The problem of column performance degradation due to irreversible binding of proteins encountered in ion chromatographic (IC) analysis of ions in protein-containing samples was overcome by using zwitterionic micelles (e.g., Zwittergent-3–14) as a portion of the eluent. A zwitterionic micellar eluent showed high ability for solubilization of proteins, and, hence, the protein-containing samples could be analyzed without need for deproteinization. On the other hand, the zwitterionic micelle was insensitive to conductivity but interacted with the analyte ions, due mainly to its unique configuration of charges (namely, the zwitterionic micelle containing both positively and negatively charged groups but carrying no net charge). Using a zwitterionic micellar eluent, the analyte ions could be detected selectively and sensitively, and moreover, the selectivity for the analyte ions was unique. A conventional anion-exchange column conditioned with a Zwittergent-3-14 micellar eluent was applied for the analysis of real biological samples (serum and urine) with direct sample injection. The results of the successful detection of inorganic anions (Cl, SO4 2–, NO2 , Br, and NO3 ) have demonstrated the usefulness of this new IC approach for the analysis of biological samples.  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