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.
Traditional micelle self‐assembly is driven by the association of hydrophobic segments of amphiphilic molecules forming distinctive core–shell nanostructures in water. Here we report a surprising chaotropic‐anion‐induced micellization of cationic ammonium‐containing block copolymers. The resulting micelle nanoparticle consists of a large number of ion pairs (≈60 000) in each hydrophobic core. Unlike chaotropic anions (e.g. ClO4?), kosmotropic anions (e.g. SO42?) were not able to induce micelle formation. A positive cooperativity was observed during micellization, for which only a three‐fold increase in ClO4? concentration was necessary for micelle formation, similar to our previously reported ultra‐pH‐responsive behavior. This unique ion‐pair‐containing micelle provides a useful model system to study the complex interplay of noncovalent interactions (e.g. electrostatic, van der Waals, and hydrophobic forces) during micelle self‐assembly.  相似文献   

2.
Recently reported ionophore‐based ion‐selective nanospheres contained pH‐independent and positively charged solvatochromic dyes. Here, we evaluate systematically the effect of anions to the fluorescence response of the nanospheres. The anion interference was found significant for anion concentrations above 10 mM. The sensor responses in the presence of various anion background was studied. While target ion (K+) causes the fluorescence of the nanospheres to decrease, increasing anion background also leads to lower fluorescence intensity. Lipophilic anions such as ClO4?, SCN?, and I? exhibited much more interference than hydrophilic anions (e. g., NO3?, Cl?, F?, SO42?). The trend of the anion interference followed the Hofmeister series. A theoretical model was also demonstrated based on anion adsorption on the surface of the nanospheres.  相似文献   

3.
A new complex of aniline (An) with tetracyanoquinodimethane (TCQM) An(TCQM)4 has been obtained and studied. The redox cycling of polyaniline (PAni-sulfo) in a neutral electrolyte containing Cl and ClO4? anions led to a replacement of the sulfate anion by the Cl and ClO4? anions. The TCQM anion can also replace the sulfate anions in polyaniline (PAni) during the cycling of the latter in a neutral electrolyte containing TCQM anions. The introduction of TCQM in PAni creates conditions for generating nanocrystalline regions inside the polymer.  相似文献   

4.
The differential capacity and the surface charge density curves as a function of the electrode potential for mercury/electrolyte solution in nitromethane interface are presented. For all the systems studied the capacity hump at the anodic potential region is observed. The height and the location of the hump considerably depends on the kind of anion. As a test of specific adsorption of ions in the systems studied the Esin-Markov effect was examined. The results indicated that anions appear to be specifically adsorbed from nitromethane in the order PF6?<ClO4?<Cl?<SCN?.  相似文献   

5.
Differential capacitance curves in the (In-Ga)/[N-methylformamide + mc KCl + (1 ? m)c KClO4] and (In-Ga)/[N-methylformamide + mc KBr + (1 ? m)c KClO4] systems are measured using an ac bridge for the following molar portions m of the surface-active anion: 0, 0.01, 0.02, 0.05, 0.1, 0.2, 0.5, and 1. The Cl? and Br? anions specific adsorption in the systems can be described quantitatively by the Frumkin isotherm. The principal parameters of Cl? and Br? anions adsorption at the (In-Ga)/N-methylformamide interface are determined by different methods. Unlike Ga/N-methylformamide interface, where the adsorption energy increased in the sequence I? ≈ Br? < Cl?, at the (In-Ga)/N-methylformamide interface it increased in the reverse sequence: Cl? < Br? < I?. The adsorption parameters at the charge density q = 0, obtained by three different methods, are close to each other. However, the parameters α1 and α2, which characterize the charge effect on the adsorption energy, when determined by the analyzing of dependences of adsorption potential drop E ads on ln(mc), differ from those determined by two other methods. The error may be caused by the assuming that the adsorption potential drop is proportional to the coverage of dense layer with the specifically adsorbed ions.  相似文献   

6.
In the structure of the title compound, C6H18N22+·H(C2H2ClO2)2·Cl, the hexane‐1,6‐diaminium dication is disordered over two sets of positions, with almost equal occupancies. Both alternative positions of the dication are in the fully extended conformation, situated on an inversion centre at (, , ). Two chloroacetic acid moieties, related by another centre of symmetry at (, , ), are connected by a very short symmetrical O...H...O hydrogen bond [O...O = 2.452 (2) Å], with the H atom at the centre of inversion. These two fragments thus effectively form the hydrogen bis(chloroacetate) monoanion, and the overall charge is balanced by an additional chloride anion which resides on a twofold axis. The ions form a layer structure, with alternating layers of dications and anions running along the [101] direction, linked via hydrogen bonds. There are two N—H...O interactions and two N—H...Cl interactions.  相似文献   

7.
A stilbene‐based compound ( 1 ) has been prepared and was highly selective for the detection of cyanide anion in aqueous media even in the presence of other anions, such as F?, Cl?, Br?, I?, ClO4?, H2PO4?, HSO4?, NO3?, and CH3CO2?. A noticeable change in the color of the solution, along with a prominent fluorescence enhancement, was observed upon the addition of cyanide. The color change was observed upon the nucleophilic addition of the cyanide anion to the electron‐deficient cyanoacrylate group of 1 . The spectral changes induced by the reaction were analyzed by comparison with two model compounds, such as compound 2 with dimethyl substituents and compound 3 without a cyanoacrylate group. An intramolecular charge‐transfer (ICT) mechanism played a key role in the sensing properties, and the mechanism was supported by DFT/TDDFT calculations.  相似文献   

8.
A series of C3i‐symmetric bicapped trigonal antiprismatic Cd8 cages [2X@Cd8L6(H2O)6] ? n Y ? solvents (X=Cl?, Y=NO3?, n=2: MOCC‐4 ; X=Br?, Y=NO3?, n=2: MOCC‐5 ; X=NO3?, Y=NO3?, n=2: MOCC‐6 ; X=NO3?, Y=BF4?, n=2: MOCC‐7 ; X=NO3?, Y=ClO4?, n=2: MOCC‐8 ; X=CO32?, n=0: MOCC‐9 ), doubly anion templated by different anions, were solvothermally synthesized by means of a flexible ligand. Interestingly, the CO32? template for MOCC‐9 was generated in situ by two‐step decomposition of DMF solvent. For other MOCCs, spherical or trigonal monovalent anions could also play the role of template in their formation. The template abilities of these anions in the formation of the cages were experimentally studied and are discussed for the first time. Anion exchange of MOCC‐8 was carried out and showed anion‐size selectivity. All of the cage‐like compounds emit strong luminescence at room temperature.  相似文献   

9.
《Analytical letters》2012,45(7):1415-1421
Abstract

The coloration reagent - leucomethylene blue, the reduction product of methylene blue, is used to determine the concentration of chlorine dioxide in the presence of Cl2 and anion species such as OCl?, ClO2 ? and ClO3 ?. This simple spectrophotometric method is performed by using the extractant 1,2-dichloroethane at pH 1.3. The linear range of ClO2 measurement extends to 0.95 mgl?1 with a detection limit of 0.02 mgl?1. The presence of chlorine and hypochlorite ion can be masked by oxalic acid. The permissible maximum concentration of ClO2 ? is 2.0 mgl?1, and ClO3 ? anion does not interfere with the measurement.  相似文献   

10.
Interactions of anions with simple aromatic compounds have received growing attention due to their relevancy in various fields. Yet, the anion–π interactions are generally very weak, for example, there is no favorable anion–π interaction for the halide anion F? on the simplest benzene surface unless the H‐atoms are substituted by the highly negatively charged F. In this article, we report a type of particularly strong anion–π interactions by investigating the adsorptions of three halide anions, that is, F?, Cl?, and Br?, on the hydrogenated‐graphene flake using the density functional theory. The anion–π interactions on the graphene flake are shown to be unexpectedly strong compared to those on simple aromatic compounds, for example, the F?‐adsorption energy is as large as 17.5 kcal/mol on a graphene flake (C84H24) and 23.5 kcal/mol in the periodic boundary condition model calculations on a graphene flake C113 (the supercell containing a F? ion and 113 carbon atoms). The unexpectedly large adsorption energies of the halide anions on the graphene flake are ascribed to the effective donor–acceptor interactions between the halide anions and the graphene flake. These findings on the presence of very strong anion–π interactions between halide ions and the graphene flake, which are disclosed for the first time, are hoped to strengthen scientific understanding of the chemical and physical characteristics of the graphene in an electrolyte solution. These favorable interactions of anions with electron‐deficient graphene flakes may be applicable to the design of a new family of neutral anion receptors and detectors. © 2012 Wiley Periodicals, Inc.  相似文献   

11.
Halide anion‐doped bismuth terephthalate hybrids were synthesized using a facile solvothermal method. Four series of hybrids doped with halide anions X? (F?, Cl?, Br? and I?) were produced by varying the molar ratios (n) of X? relative to Bi(NO3)3 (n = 0.25, 0.5, 0.75 and 1) in dimethylformamide solution. The results indicated that 0.25 equiv. of different halide anion‐doped bismuth terephthalate hybrids, especially BiBDC‐Cl(0.25) and BiBDC‐Br(0.25), exhibited excellent photocatalytic activity under visible light and UV light irradiation. They also exhibited excellent adsorption performance for Rhodamine B which could be attributed to high surface areas and negative charge on the surface of the catalysts. Moreover, the degradation of Rhodamine B under visible light irradiation is a photosensitization process and ?O2? is the most important active species. The halide anion‐doped bismuth terephthalate hybrids are promising photocatalysts for removal of organic pollutants. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

12.
The energy barrier to magnetisation relaxation in single‐molecule magnets (SMMs) proffers potential technological applications in high‐density information storage and quantum computation. Leading candidates amongst complexes of 3d metals ions are the hexametallic family of complexes of formula [Mn6O2(R‐sao)6(X)2(solvent)y] (saoH2=salicylaldoxime; X=mono‐anion; y=4–6; R=H, Me, Et, and Ph). The recent synthesis of cationic [Mn6][ClO4]2 family members, in which the coordinating X ions were replaced with non‐coordinating anions, opened the gateway to constructing families of novel [Mn6] salts in which the identity and nature of the charge balancing anions could be employed to alter the physical properties of the complex. Herein we demonstrate initial experiments to show that this is indeed possible. By replacing the diamagnetic ClO4? anions with the highly anisotropic ReIV ion in the form of [ReIVCl6]2?, the energy barrier to magnetisation relaxation is increased by up to 30 %.  相似文献   

13.
14.
The sensing mechanism of the N‐Phenyl‐N′‐(3‐quinolinyl)urea (PQU) chemosensor for fluoride anion has been investigated by density functional theory/time‐dependent density function theory. The double intermolecular hydrogen bonds are formed between the three anions (X??F?, AcO?, Cl?) and the urea fragment of PQU. In the S0 states, the Hb? X? hydrogen bonds are slightly stronger than the Ha? X? hydrogen bonds and the fluoride‐induced deprotonation occurs at the N? Hb position rather than at the N? Ha position. Consequently, the absorption peaks, including an intramolecular charge transfer transition and a ππ* transition, are significantly red‐shifted. Thermodynamic calculations confirm that the deprotonation in the ground state is favorable in energy only when excess fluoride anion exists. Along with the S0 → S1 transition, the Ha? X? hydrogen bonds strengthen and the Hb? X? hydrogen bonds weaken. However, the emission spectra of [PQU‐Hb]?, instead of [PQU‐Ha]?, are observed upon addition of fluoride anion. © 2013 Wiley Periodicals, Inc.  相似文献   

15.
A novel combination of dispersed phase polymer nanocomposite electrolyte based on PEO8‐LiClO4+ x wt % nano‐CeO2 has been investigated. A model for ion transport mechanism has been proposed to account for substantial enhancement of its electrical conductivity by ~ 2 orders of magnitude at low volume fraction of the filler reinforcement in the polymer nanocomposite films. The strength of the proposed model is based on unambiguous evidences from FTIR, TEM, and conductivity analysis. The FTIR results provide clear role of nanofiller concentration on ion–ion interaction quantified in terms of the fraction of free anion and ion‐pairs present in the nanocomposite films and its excellent correlation with conductivity versus filler concentration. The presence of asymmetry in the ν4(ClO4?) band observed at 625 cm?1 is attributed to its resolved degeneracy suggesting the presence of both uncoordinated and cation‐coordinated ClO4? anion in the matrix due to ion–ion and ion–filler interactions assisted by Lewis acid–base interaction. The enhancement in conductivity at low concentration is possibly due to direct interaction of nano‐CeO2 with both polymer host and anions resulting in the release of ionic charges. Drastic conductivity reduction at higher concentration is related to charge immobilization because of ion/ion‐pair entrapment by local clusters of filler as evidenced in TEM. © 2008 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 60–71, 2009  相似文献   

16.
17.
Here we study experimentally and by simulations the interaction of monovalent organic and inorganic anions with hydrophobic and hydrophilic colloids. In the case of hydrophobic colloids, our experiments show that charge inversion is induced by chaotropic inorganic monovalent ions but it is not induced by kosmotropic inorganic anions. For organic anions, giant charge inversion is observed at very low electrolyte concentrations. In addition, charge inversion disappears for both organic and inorganic ions when turning to hydrophilic colloids. These results provide an experimental evidence for the hydrophobic effect as the driving force for both ion specific effects and charge inversion. In the case of organic anions, our molecular dynamics (MD) simulations with full atomic detail show explicitly how the large adsorption free energies found for hydrophobic colloids are transformed into large repulsive barriers for hydrophilic colloids. Simulations confirm that solvation free energy (and hence the hydrophobic effect) is responsible for the build up of a Stern layer of adsorbed ions and charge inversion in hydrophobic colloids and it is also the mechanism preventing charge inversion in hydrophilic colloids. Overall, our experimental and simulation results suggest that the interaction of monovalent ions with interfaces is dominated by solvation thermodynamics, that is, the chaotropic/kosmotropic character of ions and the hydrophobic/hydrophilic character of surfaces.  相似文献   

18.
Combined application of cyclic voltammetry (CV) and electrochemical quartz crystal microbalance (EQCM) technique reveals a complicated interplay between the adsorption of ammonium and lower molecular weight tetraalkyl ammonium cations and desorption of Cl? anions inside carbon micropores at low surface charge densities, which results in failure of their permselectivity. Higher negative surface charge densities induce complete exclusion (desorption) of the Cl? co‐ions, which imparts purely permselective behavior on the carbon micropores. The second fundamental effect discovered herein relates to the dominant role of anion desorption (as compared to cation adsorption), that is, overwhelming failure of permselectivity extends to high negative charge densities of the electrode in the presence of bulky tetraalkyl ammonium cations, which tend to be confined in the micropores of the carbon. The results obtained are important for advancement of high power density carbon‐based supercapacitors, nanofiltration technologies with porous carbon membranes, and studies of ionic transport across biological membranes.  相似文献   

19.
The voltammetric oxidation of nickel amalgam from the hanging mercury drop electrode in aqueous solutions of F?, Cl?, Br?, I?, N3?, SCN?, and ClO4? ions have been investigated. Concentrations of these anions were sufficiently low to depress the formation of complexes with nickel(II) in the bulk of the solution.An increase in the rate of anodic oxidation with increase of concentration of anions was observed both without and with correction for the φ2 potential. This increase is explained as due to a catalytic effect of anions adsorbed on the electrode surface.Using the concept of changes of the activity coefficient of the activated complex it was possible to show that the oxidation of the nickel amalgam in thiocyanates and azides proceeds by the formation of the activated complex with bound SCN? and N3? anions. These complexes form only in the activated state and decompose when products leave the double layer.In chlorides and bromides a similar mechanism is suggested only at larger surface concentration of anions. At lower surface concentration and in iodides the oxidation proceeds by the activated complex with no anions bound to the nickel, only long-range interactions of adsorbed anions with activated complex then exist.The order of these electrode reactions was calculated using the concept of the surface activity.The two-step mechanism of the charge transfer is also discussed.  相似文献   

20.
Novel fluorescent chemosensor with good selectivity for F? anion was designed and synthesized. The sensor has a bearing on a single functionalized pillar[5]arene and Fe3+ metal complex (PN‐Fe), which showed prominent fluorescent response for F? anion over other common anions (Cl?, Br?, I?, AcO?, HSO4?, H2PO4?, ClO4?, CN? and SCN?). These results were evaluated by fluorescent method. The detection limit of PN‐Fe to F? was calculated to be 2.50×10?7 mol/L. Moreover, the sensor PN‐Fe3+ might serve as a recyclable component in sensing materials.  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