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.
The Potential step measurements are carried out on single beads of nickel hydroxide and the results are interpreted with a dual structure model featuring fast and slow diffusing components The intrinsic diffusion coefficients for the two components are found to be in the order of magnitude 10^-7 and 10^-13-10^-14 cm^2s^-1, respectively, with an apparent value for the slow component in the order of 10^-10 cm^2s^-1.  相似文献   

2.
The behavior of hydroxide and hydrated protons, the auto‐ionization products of water, at surfaces is important for a wide range of applications and disciplines. However, it is unknown at which bulk concentration these ions start to become surface active at the water–air interface. Here, we report changes in the D2O–air interface in the presence of excess D+hyd/OD?hyd determined using surface‐sensitive vibrational sum‐frequency generation (SFG) spectroscopy. The onset of the perturbation of the D2O surface occurs at a bulk concentration as low as 2.7±0.2 mm D+hyd. In contrast, a concentration of several hundred mm OD?hyd is required to change the D2O surface. The hydrated proton is thus orders of magnitude more surface‐active than hydroxide at the water–air interface.  相似文献   

3.
Anhydrous silicophosphoric acid glass with an approximate composition of H5Si2P9O29 was synthesized and its thermal and proton-conducting properties were characterized. Despite exhibiting a glass transition at 192 °C, the supercooled liquid could be handled as a solid up to 280 °C owing to its high viscosity. The glass and its melt exhibited proton conduction with a proton transport number of ∼1. Although covalent O−H bonds were weakened by relatively strong hydrogen bonding, the proton conductivity (4×10−4 S cm−1 at 276 °C) was considerably lower than that of phosphoric acid. The high viscosity of the melt was due to the tight cross-linking of phosphate ion chains by six-fold-coordinated Si atoms. The low proton conductivity was attributed to the trapping of positively charged proton carriers around anionic SiO6 units (expressed as (SiO6/2)2−) to compensate for the negative charges.  相似文献   

4.
The effect of ionizing radiation on DNA constituents is a widely studied fundamental process using experimental and computational techniques. In particular, radiation effects on nucleobases are usually tackled by mass spectrometry in which the nucleobase is embedded in a water nanodroplet. Here, we present a multiscale theoretical study revealing the effects and the dynamics of water droplets towards neutral and ionized thymine. In particular, by using both hybrid quantum mechanics/molecular mechanics and full ab initio molecular dynamics, we reveal an unexpected proton transfer from thymine cation to a nearby water molecule. This leads to the formation of a neutral radical thymine and a Zundel structure, while the hydrated proton localizes at the interface between the deprotonated thymine and the water droplet. This observation opens entirely novel perspectives concerning the reactivity and further fragmentation of ionized nucleobases.  相似文献   

5.
The absolute solvation energies (free energies and enthalpies) of the proton in ammonia are used to compute the pKa of species embedded in ammonia. They are also used to compute the solvation energies of other ions in ammonia. Despite their importance, it is not possible to determine experimentally the solvation energies of the proton in a given solvent. We propose in this work a direct approach to compute the solvation energies of the proton in ammonia from large-sized neutral and protonated ammonia clusters. To undertake this investigation, we performed a geometry optimization of neutral and protonated ammonia 30-mer, 40-mer, and 50 mer to locate stable structures. These structures have been fully optimized at both APFD/6-31++g(d,p) and M06-2X/6-31++g(d,p) levels of theory. An infrared spectroscopic study of these structures has been provided to assess the reliability of our investigation. Using these structures, we have computed the absolute solvation free energy and the absolute solvation enthalpy of the proton in ammonia. It comes out that the absolute solvation free energy of the proton in ammonia is calculated to be −1192 kJ mol–1, whereas the absolute solvation enthalpy is evaluated to be −1214 kJ mol–1. © 2019 Wiley Periodicals, Inc.  相似文献   

6.
Dynamic protonation equilibria in water of one 4‐methylimidazole molecule as well as for pairs and groups consisting of 4‐methylimidazole, acetic acid and bridging water molecules are studied using Q‐HOP molecular dynamics simulation. We find a qualitatively different protonation behavior of 4‐methylimidazole compared to that of acetic acid. On one hand, deprotonated, neutral 4‐methylimidazole cannot as easily attract a freely diffusing extra proton from solution. Once the proton is bound, however, it remains tightly bound on a time scale of tens of nanoseconds. In a linear chain composed of acetic acid, a separating water molecule and 4‐methylimidazole, an excess proton is equally shared between 4‐methylimidazole and water. When a water molecule is linearly placed between two acetic acid molecules, the excess proton is always found on the central water. On the other hand, an excess proton in a 4‐methylimidazole‐water‐4‐methylimidazole chain is always localized on one of the two 4‐methylimidazoles. These findings are of interest to the discussion of proton transfer along chains of amino acids and water molecules in biomolecules.  相似文献   

7.
8.
《Analytical letters》2012,45(19-20):1917-1937
Abstract

Longitudinal relaxation rates of water protons were determined for aqueous solutions of orthophosphoric acid at concentrations up to 85% by weight, and their concentration dependence was investigated. An appropriate choice of variables yielded an essentially linear plot for concentrations up to at least 15% by weight. Deviations from linearity are discussed and are interpreted in terms of mass-law effects. The use of parameters of this plot for deriving values for correlation coefficients and for a hydration parameter is illustrated by application to observed data and data from the literature.  相似文献   

9.
Published data on the mechanisms of hydrogen ion transport in solids and aqueous solutions are described systematically, including defect formation, rotational mobility of proton-containing groups, proton hopping along a hydrogen bond, proton translational mobility, and proton conduction. Resorting to the authors" theoretical results and published data, the main criteria for the selection of systems possessing high proton mobility are formulated.  相似文献   

10.
The present paper discusses the behaviour of transfer free energy of some amino acids from water to 4M, 6M and 8M aqueous urea. Dissection of transfer free energy into cavity term, interaction term and electrical term reveals that cavity forming free energy of transfer ΔG0t (cav) plays an important role in dictating actual interaction of amino acids in aqueous urea. Cavity forming free energy of transfer has been estimated by using Scaled Particle Theory (SPT).  相似文献   

11.
12.
Herein, we present the comparison of a large set of experimentally measured proton affinity (PA) values for 65 aromatic carboxylate anions with the values calculated by using selected popular DFT (B3LYP, PBE0, and M05‐2X) and composite [G3(MP2), G4(MP2)] quantum chemistry methods. The root‐mean‐square error (RMSE) values for the chosen methods are RMSEPBE0=1.7, RMSEB3LYP=4.6, RMSEM05‐2X=6.6, RMSEG3MP2=6.3, RMSEG4MP2=4.5 kJ mol?1. In the second part of the study, 82 PA values for substituted phenide ions and a few heteroaromatic anions were calculated. Again, very good agreement between the calculated and experimental values has been observed: RMSEPBE0=1.9, RMSEB3LYP=4.5, RMSEM05‐2X=6.3, RMSEG3MP2=4.9, RMSEG4MP2=5.5 kJ mol?1. Our results show that, for medium‐sized carboxylate anions, all tested methods give reliable results and, surprisingly, much more computationally demanding composite methods do not perform significantly better than the time‐efficient DFT methods.  相似文献   

13.
Present paper discusses the behaviour of transfer free energy of some amino acids from water to aqueous solution of Ethanol, 2‐PrOH ad t‐BuOH at different compositions. Dissection of transfer free energy into cavity term, interaction term and electrical term reveals that cavity forming free energy of transfer ΔGt0 (Cav.) plays an important role in dictating actual interaction of amino acids in these mixed solvents. Cavity forming free energy of transfer has been estimated by using Scaled Particle Theory (SPT).  相似文献   

14.
Imaging Ca2+ dynamics in living systems holds great potential to advance neuroscience and cellular biology. G‐GECO1.1 is an intensiometric fluorescent protein Ca2+ biosensor with a Thr‐Tyr‐Gly chromophore. The protonated chromophore emits green upon photoexcitation via excited‐state proton transfer (ESPT). Upon Ca2+ binding, a significant population of the chromophores becomes deprotonated. It remains elusive how the chromophore structurally evolves prior to and during ESPT, and how it is affected by Ca2+. We use femtosecond stimulated Raman spectroscopy to dissect ESPT in both the Ca2+‐free and bound states. The protein chromophores exhibit a sub‐200 fs vibrational frequency shift due to coherent small‐scale proton motions. After wavepackets move out of the Franck–Condon region, ESPT gets faster in the Ca2+‐bound protein, indicative of the formation of a more hydrophilic environment. These results reveal the governing structure–function relationship of Ca2+‐sensing protein biosensors.  相似文献   

15.
The effect of cucurbit[7]uril (CB[7]) nano‐caging on the photophysical properties, particularly excited‐state proton transfer (ESPT) reaction, of an eminent anti‐cancer drug, topotecan (TPT), is demonstrated through steady‐state and time‐resolved fluorescence measurements. TPT in water (pH 6) exists exclusively as the cationic form (C) in the ground state. However, the drug emission mainly comes from the excited‐state zwitterionic form (Z*) of TPT, and is attributed to water‐assisted ESPT between the 10‐hydroxyl group and water, which leads to the transformation of C* to Z* of TPT. In the presence of CB[7], it is found that selective encapsulation of the C form of TPT results in the formation of a 1:1 inclusion complex (CB[7]:TPT), and the ESPT process is inhibited by this encapsulation process. As a result, C* becomes the dominant emitting species in the presence of CB[7] rather than Z*, and fluorescence switching takes place from green to blue. Time‐resolved studies also support the existence of CB[7]‐encapsulated cationic species as the major emitting species in the presence of the macrocyclic host. Semi‐empirical quantum chemical calculations are employed to gain insight into the molecular picture of orientation of TPT in the inclusion complex. It is clearly seen from the optimised structure of 1:1 CB[7]:TPT inclusion complex that both 10‐hydroxyl and 9‐dimethylaminomethylene groups of TPT lie partly inside the cavity, and thereby inhibit the excited‐state transformation of C* to Z* by the ESPT process. Finally, controlled release of the drug is achieved by means of fluorescence switching by introducing NaCl, which is rich in cells, as an external stimulus.  相似文献   

16.
17.
18.
Photoacids attract increasing scientific attention, as they are valuable tools to spatiotemporally control proton-release reactions and pH values of solutions. We present the first time-resolved spectroscopic study of the excited state and proton-release dynamics of prominent merocyanine representatives. Femtosecond transient absorption measurements of a pyridine merocyanine with two distinct protonation sites revealed dissimilar proton-release mechanisms: one site acts as a photoacid generator as its pKa value is modulated in the ground state after photoisomerization, while the other functions as an excited state photoacid which releases its proton within 1.1 ps. With a pKa drop of 8.7 units to −5.5 upon excitation, the latter phenolic site is regarded a super-photoacid. The 6-nitro derivative exhibits only a phenolic site with similar, yet slightly less photoacidic characteristics and both compounds transfer their proton to methanol and ethanol. In contrast, for the related 6,8-dinitro compound an intramolecular proton transfer to the ortho-nitro group is suggested that is involved in a rapid relaxation into the ground state.  相似文献   

19.
Alcohols, the simplest amphiprotic organic compounds, can exhibit either acidic or basic behavior by donating or accepting a proton. In this study, proton dissociation of a model photoacid in solution is explored by using time‐resolved spectroscopy, revealing quantitatively for the first time that alcohol acts as a Brønsted base because of H‐bonded cluster formation to enhance the reactivity. The protonated alcohol cluster, the alkyl oxonium ion, can be regarded as a key reaction intermediate in the well‐established alcohol dehydration reaction. This finding signifies, as in water, the cooperativity of protic solvent molecules to facilitate nonaqueous acid–base reactions.  相似文献   

20.
Standard non‐semiempirical continuum‐dielectric orbital‐based methods horribly overpredict, by 26‐50 kcal mol−1, the Gibbs energy for the water autoionization reaction 2 H2O(l) → H3O+(aq) + OH(aq). Here, we demonstrate these errors, fully investigate the reasons for these errors, and show that the use of 4 explicit solvent within the continuum (the “semicontinuum,” “cluster‐continuum,” or “hybrid” technique) can reduce the error of a standard continuum model from 50 to 2 kcal mol−1. Results from pure cluster, pure continuum (several versions including semiempirical ones), and semicontinuum modeling are each presented and discussed. We recommend use of 3 waters around hydronium and 4 waters around hydroxide with standard continua whenever these ions are involved in reaction. To the possible surprise of some, time‐consuming molecular‐dynamics simulations are not needed to reproduce this problematic energy.  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