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 187 毫秒
1.
The dielectric properties of aqueous solutions of ammonia, hydrogen fluoride, hydrogen peroxide, and ammonium halides are determined by the number of hydrogen bonds formed by the solute species with the molecules forming the water framework. For such solutions, the product of the concentration coefficient of the dielectric permittivity by the ratio of the number of hydrogen bonds per solute molecule and per water molecule is constant. In aqueous solutions with a water-like structure, the mechanism of the dielectric polarization is the same as in pure water (proton-activation mechanism) and is determined by proton fluctuations on the line of hydrogen bonds.__________Translated from Zhurnal Obshchei Khimii, Vol. 75, No. 1, 2005, pp. 39–45.Original Russian Text Copyright © 2005 by Potapov, Parkhomenko.  相似文献   

2.
An electrochemical method for proton transport visualization was developed and applied to the investigation of proton-conducting membrane materials. The method employs the change in the visual appearance of chemo-chromic tungsten oxide WO3 in the presence of atomic hydrogen. An all-solid electrochemical cell arranged by substituting a fuel cell cathode with a thin film of WO3-electrode was built and shown to generate both optical and electrical response to hydrogen gas exposure. The design of the cell was extended to a high throughput screening system that was utilized to characterize proton transport properties of samples, including a number of new compounds synthesized in-house by sol–gel wet chemistry. Non-destructive introduction of superacidic groups promoting proton hopping in the membrane materials was achieved by photodecomposition of a photoacid generator just after membrane casting. A model quantitatively describing current–voltage relation in the cell was developed and successfully applied to derive area-specific resistance of proton-conductive membranes from the experimental results. Area-specific resistances of membranes are derived from the slopes of optically reconstructed voltage–current curves. Sensitivity and dynamic range of the screening method are discussed.  相似文献   

3.
A huge dielectric effect has been observed in a pure and water-soluble hydrogen-bonded organic crystal, 1,4-diazabicyclo[2.2.2]octane hydroiodide [C6H13N2]+.I(-) (dabcoHI). In this structure, the dabco cations are NH+...N bonded into linear aggregates, where the protons are disordered at two nitrogen atoms and the crystal acquires the symmetry of space group P6m2. This nonpolar crystal exhibits a barely temperature-dependent dielectric constant exceeding 1000 at ambient conditions. The dielectric response is extremely anisotropic, more than 2 orders of magnitude higher along the linear hydrogen bonded chains than in perpendicular directions. The physics underlying this effect originates from proton transfers in the NH+...N bonds, leading to disproportionation defects and formation of polar nanodomains, which, on the macroscopic scale, results in one-dimensional relaxor ferroelectricity. Such properties are unprecedented for the materials with hydrogen bonds highly polarizable due to proton disorder. The proton disordering in dabcoHI is analogous to this in H2O ice, where the hydrogen bonds remain disordered until the lowest temperature.  相似文献   

4.
Computer simulations of ice Ih with different proton orientations are presented. Simulations of proton disordered ice are carried out using a Monte Carlo method which samples over proton degree of freedom, allowing for the calculation of the dielectric constant and for the examination of the degree of proton disorder. Simulations are also presented for two proton ordered structures of ice Ih, the ferroelectric Cmc2(1) structure or ice XI and the antiferroelectric Pna2(1) structure. These simulations indicate that a transition to a proton ordered phase occurs at low temperatures (below 80 K). The symmetry of the ordered phase is found to be dependent on the water potential. The stability of the two proton ordered structures is due to a balance of short-ranged interactions which tend to stabilize the Pna2(1) structure and longer-range interactions which stabilize the Cmc2(1) structure.  相似文献   

5.
According to the DFT calculations (B3LYP/6-31G*), the structure of solvate complexes formed by N-methyltrifluoromethanesulfonamide with phosgene, formaldehyde, and DMSO and the mode of hydrogen bonding therein (two-center or bifurcate three-center) depend on the proton affinity of the onium base and composition of the complex. The 1 : 1 cyclic dimer–phosgene complex and 1 : 2 cyclic dimer–formaldehyde complex are stabilized by one or two bifurcated hydrogen bonds, respectively.  相似文献   

6.
The effect of solvent polarity versus specific C–HO contacts on the vibrational νC–H mode is studied using CHCl3 as a model system. Ab initio SCI–PCM calculations show that the overall shift of the νC–H band, sometimes ascribed to the C–HO hydrogen bonding, can in fact be explained by the electrostatic interaction with a dielectric environment. The presence of a new νC–H band – assigned to the C–HO bonded forms – remains as the most reliable evidence of C–HO hydrogen bonding.  相似文献   

7.
Summary The structure of hydrofluorides with the composition MeF2 · HF(Me=Sr, Ba) includes two anions: F and HF2 with a linear hydrogen bond. In SrF2 · HF, the hydrogen bond is characterized by a more asymmetrical position of the proton than in BaF2 · HF.Increasing the number of added HF molecules leads to the formation of two bifluoride ions in the compounds CaF2 · 2HF and SrF2 · 2.5HF. The hydrogen bond in this case is also nonsymmetrical, and the average F–H distance is equal to 1.17 A. Thus, the hydrogen bond in these compounds is of the same order as in KF · 2HF (rF–F=2.33 A) [12],i.e., weaker than in KF · HF (rF–F=2.26 A) [13], but stronger than in solid HF (rF–F=2.49 A) [14].A structure with two polyfluoride anions evidently corresponds to the compound BaF2 · 3HF.Institute of Inorganic Chemistry, Siberian Branch of the Academy of Sciences of the USSR. Translated from Zhurnal Strukturnoi Khimii, Vol. 9, No. 2, pp. 202–206, March–April, 1968.  相似文献   

8.
A study has been made of hydrogen bonding in two pairs of proton donors: 2,7-dihydroxynaphthalene and 2-hydroxynaphthalene, and malononitrile and methylmalononitrile. In each pair one of the proton donors contains one reactive center, and the other two. It is shown by NMR and quantum-chemical methods that molecules possessing two equivalent proton donor centers form hydrogen bonds involving both protons, the formation of one hydrogen bond having practically no effect on the other.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 26, No. 3, pp. 340–344, May–June, 1990.  相似文献   

9.
Calculations have been carried out to optimize the structure of Van der Waals complexes of methanol with N-methyl-2-nitroaniline, a dye capable of shifts in visible and ultraviolet spectra that depend on (1) solvent dielectric, (2) solvent shell structure, and (3) hydrogen bonding to a slight extent. Hartree–Fock–Roothaan calculations with various basis sets and single-excitation configuration interaction (SCI) are compared to Density–Functional–Theory Time-Dependent Hartree–Fock (DFT-TD) results for three low-energy ultraviolet electronic transitions. Energy-minimized structures are reported for a trimeric complex of two methanol-one water as found using a 6-311G** basis indicating two possible hydrogen-bonding schemes. The effect of a dielectric medium on the ultraviolet spectrum is compared to gas-phase clusters. Electronic transitions are also given for the dye-probe complexed with four or five methanol molecules finding good agreement with observed shifts in the ultraviolet spectrum as found with the TDHF-DFT formalism for the lowest energy transition near 425nm.  相似文献   

10.
The structure of Ba2In2O4(OH)2 is analysed by the explicit full optimization of a large number of possible proton arrangements using periodic density functional theory. It is shown that the experimental assignments in which protons appear to be located at high symmetry positions with unphysical bond lengths do not correspond to minima on the potential energy hypersurface. The apparent sites are averages of a number of possible proton locations involving a set of possible local structural environments in which the internuclear separations are more realistic. Such problems with structural refinements are common where profile refinement programs place the atoms at the average position due to dynamic and/or static disorder. Thus while the calculations support a previous neutron diffraction analysis of the structure in that the average structure contains two different proton sites, they also reveal substantial information about the local environments of the protons. In all optimizations, the protons moved from the average positions suggested in the neutron diffraction study with calculated O–H and OHO distances consistent with those observed in other oxides. The energies of different proton distributions vary significantly so the protons are not randomly distributed. We also present an analysis of the vibrational properties of the O–H bonds. Since the strength of the hydrogen bonds is closely related to the local structural environments of the protons, a range of vibrational frequencies is obtained providing a prediction of the vibrational spectra. In O–HO linkages, O–H stretching modes soften with increasing HO hydrogen bond strength, while the in-plane and out-of-plane bending or libration modes stiffen. Together, our results show how modern theoretical methods can provide a clearer understanding of the structure and dynamics of a complex inorganic material.  相似文献   

11.
Main transport properties were studied for selective membranes with low dielectric constants based on liquid ion exchangers involving nitrogen-containing organic base cations. Permeabilities and ion flows through a membrane were calculated for major and interfering ions. Dependences of the transport properties of membranes on the concentrations of the ion exchanger and near-membrane solution and their potentiometric characteristics are presented. It was demonstrated that the transport properties of liquid membranes are determined by two main factors: the transfer of counterions through the phase boundary by the extraction–exchange mechanism and the leaching of the ion exchanger from the membrane.  相似文献   

12.
SrTiO3 powders were prepared using a sol–gel process. The electrical properties and the microstructure of doped SrTiO3-based capacitor-varistor double functions ceramics prepared by the sol–gel process were compared with those of ceramics prepared via the solid-state reaction method using X-ray diffraction, scanning electron microscopy, and measurements of dielectric constants and current-voltage characteristics. The sol–gel based materials exhibited enhanced capacitor-varistor properties compared to those prepared by a conventional solid-state method. These enhanced properties are attributed to the smaller, more uniform grain structure obtained in the sol–gel-derived material, which promotes the formation of an effective core-shell structure in which an acceptor-doped layer encapsulate the donor-doped SrTiO3 grains. The effect of such microstructures on electrical properties is discussed.  相似文献   

13.
The electronic structure of the model hydrogen-bonded systems has been studied at the all-valence level in relation to the charge transfer mechanism. It is concluded that together with the intermolecular proton transfer an electronic charge transport occurs for hydrogen bonds of 2.70–3.00 Å in length, i.e., when the proton motion within the bond is anticipated. For elucidation of transport properties of one-dimensional hydrogen-bonded systems the tunneling–hopping model is preferred instead of the band theory. The importance of the proposed mechanism of the charge transfer for biological processes has been emphasized.  相似文献   

14.
The kinetics and mechanism of proton exchange between thiocarboxylic acids and o-substituted phenols with intramolecular hydrogen bonds (IMHB) were studied by dynamic PMR spectroscopy. Fast uncatalyzed exchange in cyclic binary complexes is observed for phenols with weak IMHB. In systems with strong IMHB fast exchange occurs only in the presence of an alkaline catalyst, the reaction of which with the phenol leads to cleavage of the IMHB. Concerted transfer of two protons and a cation is realized in the final step of the reaction. According to the proposed mechanism, the measured energy of activation of proton exchange is the sum of the energies of activation of the steps involving cleavage of the IMHB and strictly proton transfer.Translated from Teoreticheskaya i éksperimental'naya Khimiya, Vol. 24, No. 1, pp. 49–56, January–February, 1988.  相似文献   

15.
We discuss the minimum size of ice nanoparticles in water-rich glycerol-water mixtures, as studied by broadband dielectric spectroscopy (BDS) in the frequency range from 1 Hz to 250 MHz and differential scanning calorimetry (DSC) at the temperature interval from 138 to 313 K. It is known that the extra water which is free from the glycerol hydrogen bond network forms the water cooperative domain. This cooperative domain leads to a freezing of water. With the formation of the frozen water state, another distinct water structure forms on the interface between the ice nanocrystal and mesoscopic glycerol-water domain. The mole fractions of different stages of water (i.e., water molecules in the mesoscopic domain, ice nanocrystals, and the interface between the two) were determined, and the minimum number of water molecules that can gain the bulk ice properties was estimated as approximately 300 water molecules.  相似文献   

16.
Complex reflection coefficients for 2-ethoxyethanol–dimethylformamide (DMF), ethanol–DMF, and 1-propanol–DMF mixtures at several temperatures from 20 to 50° and the frequency range 10 MHz to 10 GHz were determined by time-domain spectroscopy in reflection mode. Fourier transforms and least-squares fitting were used to obtain complex permittivity, static dielectric constant, and relaxation time. The excess dielectric parameters, Kirkwood correlation factors, and thermodynamic properties for the binary mixtures were also determined. The static dielectric constant for the mixtures was fitted well with the modified Bruggeman model.  相似文献   

17.
A new mechanism of proton transfer in donor–acceptor complexes with long hydrogen bonds is suggested. The transition is regarded as totally adiabatic. Two closest water molecules that move synchronously by hindered translation to and from the reaction complex are crucial. The water molecules induce a shift of the proton from the donor to the acceptor with simultaneous breaking/formation of hydrogen bonds between these molecules and the proton donor and acceptor. Expressions for the activation barrier and kinetic hydrogen isotope effect are derived. The general scheme is illustrated with the use of model molecular potentials, and with reference to the excess proton conductivity in aqueous solution.  相似文献   

18.
The mechanisms of the proton transfer in associates of two molecules of formic acid with one molecule of hydrazine or hydrogen peroxide were studied usingab initio (SCFj6-31G**) method. The mechanism of cooperative (concerted, one-step) four-proton transfer is realized in the associate with the hydrazine molecule. The proton transfer occurs stepwisevia an intermediate in the associate with a hydrogen peroxide molecule. The calculated activation barriers to the proton transfer in the associates investigated are 34.7 kcal mol–1 and 27.1 kcal mol–1, respectively.Translated fromlzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 11, pp. 2631–2635, November, 1996.  相似文献   

19.
Ab initio calculations at second-order Møller-Plesset perturbation theory with the 6-31 + G(d,p) basis set have been performed to determine the equilibrium structures and energies of a series of negative-ion hydrogen-bonded complexes with H2O, H2S, HCN, and HCl as proton donors and OH, SH, CN, and Cl as proton acceptors. The computed stabilization enthalpies of these complexes are in agreement to within the experimental error of 1 kcal mol–1 with the gas-phase hydrogen bond enthalpies, except for HOHOH, in which case the difference is 1.8 kcal mol–1. The structures of these complexes exhibit linear hydrogen bonds and directed lone pairs of electrons except for complexes with H2O as the proton donor, in which cases the hydrogen bonds deviate slightly from linearity. All of the complexes have equilibrium structures in which the hydrogen-bonded proton is nonsymmetrically bound, although the symmetric structures of HOHOH and ClHCl are only slightly less bound than the equilibrium structures. MP2/6-31 + G(d,p) hydrogen bond energies calculated at optimized MP2/B-31 + G(d,p) and at optimized HF/6-31G(d) geometries are similar. Using HF/6-31G(d) frequencies to evaluate zero-point and thermal vibrational energies does not introduce significant error into the computed hydrogen bond enthalpies of these complexes provided that the hydrogen-bonded proton is definitely nonsymmetrically bound at both Hartree-Fock and MP2.  相似文献   

20.
The alkaloid colchicine forms, in addition to the previously known dihydrate host–guest complex, a monohydrate complex. The crystal structure of the monohydrate was determined by direct methods and refined to a final R value of 0.046 for 1425 observed reflections. Crystal data are: orthorhombic, space group P2 12 12 1, a = 9.145(2) Å; b = 13.270(3) Å; c = 17.942(4) Å, V = 2177(1) Å3, Z= 4, Dx = 1.22 g cm-3, T = 293 K. The conformation of the molecule is practically identical with the conformation in the dihydrate inclusion complex. Water molecules show proton donor as well as proton acceptor properties and are hydrogen bonded with the three colchicine molecules giving rise to the three dimensional H-bonded network.  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