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 0 毫秒
1.
Ensemble infrared photodissociation (IRPD) spectra in the hydrogen stretch region (~2800-3800 cm(-1)) are reported for aqueous nanodrops containing ~250 water molecules and either SO(4)(2-), I(-), Na(+), Ca(2+), or La(3+) at 133 K. Each spectrum has a broad feature in the bonded-OH region (~2800-3500 cm(-1)) and a sharp feature near 3700 cm(-1), corresponding to the free-OH stretch of surface water molecules that accept two hydrogen bonds and donate one hydrogen bond (AAD water molecules). A much weaker band corresponding to AD surface water molecules is observed for all ions except SO(4)(2-). The frequencies of the AAD free-OH stretch red-shift with increasingly positive charge, consistent with a Stark effect as a result of the ion's electric field at the droplet surface, and from which the corresponding frequency for water molecules at the surface of neutral nanodrops of this size is estimated to be 3699.3-3700.1 cm(-1). The intensity of the AAD band increases with increasing positive charge, consistent with a greater population of AAD water molecules for the more positively charged nanodrops. The spectra of M(H(2)O)(~250), M = Na(+) and I(-), are very similar, whereas those for Ca(2+) and SO(4)(2-) have distinct differences. These results indicate that the monovalent ions do not affect the hydrogen-bonding network of the majority of water molecules whereas this network is significantly affected in nanodrops containing the multivalent ions. The ion-induced effect on water structure propagates all the way to the surface of the nanodrops, which is located more than 1 nm from the ion.  相似文献   

2.
The effects of water on ion fluorescence were investigated, and average sequential water molecule binding energies to hydrated ions, M(z)(H(2)O)(n), at large cluster size were measured using ion nanocalorimetry. Upon 248-nm excitation, nanodrops with ~25 or more water molecules that contain either rhodamine 590(+), rhodamine 640(+), or Ce(3+) emit a photon with average energies of approximately 548, 590, and 348 nm, respectively. These values are very close to the emission maxima of the corresponding ions in solution, indicating that the photophysical properties of these ions in the nanodrops approach those of the fully hydrated ions at relatively small cluster size. As occurs in solution, these ions in nanodrops with 8 or more water molecules fluoresce with a quantum yield of ~1. Ce(3+) containing nanodrops that also contain OH(-) fluoresce, whereas those with NO(3)(-) do not. This indirect fluorescence detection method has the advantages of high sensitivity, and both the size of the nanodrops as well as their constituents can be carefully controlled. For ions that do not fluoresce in solution, such as protonated tryptophan, full internal conversion of the absorbed 248-nm photon occurs, and the average sequential water molecule binding energies to the hydrated ions can be accurately obtained at large cluster sizes. The average sequential water molecule binding energies for TrpH(+)(H(2)O)(n) and a doubly protonated tripeptide, [KYK + 2H](2+)(H(2)O)(n), approach asymptotic values of ~9.3 (n ≥ 11) and ~10.0 kcal/mol (n ≥ 25), respectively, consistent with a liquidlike structure of water in these nanodrops.  相似文献   

3.
Mechanistic information about how gaseous ions are formed from charged droplets has been difficult to establish because direct observation of nanodrops in a size range relevant to gaseous macromolecular ion formation by optical or traditional mass spectrometry methods is challenging owing to their small size and heterogeneity. Here, the mass and charge of individual aqueous nanodrops between 1–10 MDa (15–32 nm diameter) with ∼50–300 charges are dynamically monitored for 1 s using charge detection mass spectrometry. Discrete losses of minimally solvated singly charged ions occur, marking the first direct observation of ion emission from aqueous nanodrops in late stages of droplet evaporation relevant to macromolecular ion formation in native mass spectrometry. Nanodrop charge depends on the identity of constituent ions, with pure water nanodrops charged slightly above the Rayleigh limit and aqueous solutions containing alkali metal ions charged progressively below the Rayleigh limit with increasing cation size. MS2 capsid ions (∼3.5 MDa; ∼27 nm diameter) are more highly charged from aqueous ammonium acetate than from its biochemically preferred, 100 mM NaCl/10 mM Na phosphate solution, consistent with ion emission reducing the nanodrop and resulting capsid charge. The extent of charging indicates that the capsid partially collapses inside the nanodrops prior to the charging and formation of the dehydrated gaseous ions. These results demonstrate that ion emission can affect macromolecular charging and that conformational changes to macromolecular structure can occur in nanodrops prior to the formation of naked gaseous ions.

Ion evaporation from aqueous nanodrops is measured for the first time using charge detection mass spectrometry, and the origin of solute ion dependent charging of large (MDa) macromolecules is revealed.  相似文献   

4.
5.
6.
7.
We present a new polarizable force field for aqueous ions (Li(+), Na(+), K(+), Rb(+), Cs(+), Mg(2 +), Ca(2 +), Sr(2 +), and Cl(-)) derived from condensed phase ab initio calculations. We use maximally localized Wannier functions together with a generalized force and dipole-matching procedure to determine the whole set of parameters. Experimental data are then used only for validation purposes and a good agreement is obtained for structural, dynamic, and thermodynamic properties. The same procedure applied to crystalline phases allows to parametrize the interaction between cations and the chloride anion. Finally, we illustrate the good transferability of the force field to other thermodynamic conditions by investigating concentrated solutions.  相似文献   

8.
This work discusses pulsed electric fields (PEF) induced effects in treatment of aqueous suspensions of concentrated yeast cells (S. cerevisiae). The PEF treatment was done using pulses of near-rectangular shape, electric field strength was within E=2-5 kV/cm and the total time of treatment was t(PEF)=10(-4)-0.1 s. The concentration of aqueous yeast suspensions was in the interval of C(Y)=0-22 (wt%), where 1% concentration corresponds to the cellular density of 2x10(8) cells/mL. Triton X-100 was used for studying non-ionic surfactant additive effects. The electric current peak value I was measured during each pulse application, and from these data the electrical conductivity sigma was estimated. The PEF-induced damage results in increase of sigma with t(PEF) increasing and attains its saturation level sigma approximately sigma(max) at long time of PEF treatment. The value of sigma(max) reflects the efficiency of damage. The reduced efficiency of damage at suspension volume concentration higher than phi(Y) approximately 32 vol% is explained by the percolation phenomenon in the randomly packed suspension of near-spherical cells. The higher cytoplasmic ions leakage was observed in presence of surfactant. Experiments were carried out in the static and continuous flow treatment chambers in order to reveal the effects of mixing in PEF-treatment efficiency. A noticeable aggregation of the yeast cells was observed in the static flow chamber during the PEF treatment, while aggregation was not so pronounced in the continuous flow chamber. The nature of the enhanced aggregation under the PEF treatment was revealed by the zeta-potential measurements: these data demonstrate different zeta-potential signs for alive and dead cells. The effect of the electric field strength on the PEF-induced extraction of the intracellular components of S. cerevisiae is discussed.  相似文献   

9.
10.
In solution, half-cell potentials are measured relative to those of other half cells, thereby establishing a ladder of thermochemical values that are referenced to the standard hydrogen electrode (SHE), which is arbitrarily assigned a value of exactly 0 V. Although there has been considerable interest in, and efforts toward, establishing an absolute electrochemical half-cell potential in solution, there is no general consensus regarding the best approach to obtain this value. Here, ion-electron recombination energies resulting from electron capture by gas-phase nanodrops containing individual [M(NH3)6]3+, M = Ru, Co, Os, Cr, and Ir, and Cu2+ ions are obtained from the number of water molecules that are lost from the reduced precursors. These experimental data combined with nanodrop solvation energies estimated from Born theory and solution-phase entropies estimated from limited experimental data provide absolute reduction energies for these redox couples in bulk aqueous solution. A key advantage of this approach is that solvent effects well past two solvent shells, that are difficult to model accurately, are included in these experimental measurements. By evaluating these data relative to known solution-phase reduction potentials, an absolute value for the SHE of 4.2 +/- 0.4 V versus a free electron is obtained. Although not achieved here, the uncertainty of this method could potentially be reduced to below 0.1 V, making this an attractive method for establishing an absolute electrochemical scale that bridges solution and gas-phase redox chemistry.  相似文献   

11.
Structural effects in the radioluminescence spectra of glassy and polycrystalline aqueous solutions of chloride and alkaline ices observed by pulse radiolysis at temperatures 6–110 K are reported. The luminescence efficiency as well as position of λmax of the emission spectra are dependent on the physical state and temperature of the matrix. For all the investigated aqueous polycrystalline matrices and for H2O ice, the emission band peaking at about 330 nm, assigned to OH*A2Σ→X2Π transition appears at temperatures below 40 K. This UV band was not observed for glassy matrices. Luminescence bands observable in the visible range of spectrum (400–600 nm) can be associated with the emission of (OH)*, the radiative capture of trapped electrons et by metal cations Me+(Me2+) and trapped atoms H√t. For polycrystalline chloride matrices a contribution of the emission of (Cl2−2)* must be taken into account.  相似文献   

12.
The radial distribution functions (RDFs) of aqueous solutions of ytterbium chloride were calculated using previously collected X-ray diffraction data for a wide range of solution concentrations under ambient conditions. Different solution structure models were constructed. Theoretical RDFs were calculated for each model. The optimum models for all systems studied were chosen based on the best fit between the theoretical and experimental RDFs. Quantitative characteristics of the nearest environment of Yb3+ and Cl ions in solutions, namely, the coordination number, interparticle distances, and types of ion pairs were determined. The average number of water molecules in the first coordination sphere of the cation decreases from 8.2 to 6.0 as the solution concentration increases. The structure of all systems investigated is governed by non-contact ion associates throughout the concentration range studied.  相似文献   

13.
Literature data and own experimental results on structural characteristics of the halide ions nearest surrounding in the aqueous electrolyte solutions under standard conditions has been summarized. Structural parameters like coordination number, interparticle distance, and ion association type have been discussed. It has been shown that in the halide ions series, from fluoride to iodide, the coordination number gradually increases from 6 to 8. In the same row, the coordination sphere stability decreases, this is reflected in more asymmetrical arrangement of water molecules of the nearest surrounding.  相似文献   

14.
Addition of ligands to [Pd(η3-RCHCHCH2)(μ-Cl)]2 or chloride ions to cationic [(η3-RCHCHCH2)PdL2]+BF4 induces the formation of neutral complexes η1-RCHCHCH2PdClL2 (R=H with L=(4-ClC6H4)3P, (4-CH3C6H4)3P, (4-CF3C6H4)3P or L2=1,2-bis(diphenylphosphino)butane (dppb), 1,1′-bis(diphenylphosphino)ferrocene (dppf); R=Ph with L=(4-ClC6H4)3P), instead of the expected cationic complexes [(η3-RCHCHCH2)PdL2]+Cl. In the presence of chloride ions, the reaction of morpholine with the cationic complexes [(η3-allyl)Pd(PAr3)2]+BF4 (Ar=4-ClC6H4, 4-CH3C6H4) goes slower and involves both cationic [(η3-allyl)Pd(PAr3)2]+ and neutral η1-allyl-PdCl(PAr3)2 complexes as reactive species in equilibrium with Cl. The cationic complex is more reactive than the neutral one. However, their relative contribution in the reaction strongly depends on the chloride concentration, which controls their relative concentration. The neutral η1-allyl-PdCl(PAr3)2 may become the major reactive species at high chloride concentration. Consequently, [Pd(η3-allyl)(μ-Cl)]2 associated with ligands or cationic [(η3-allyl)PdL2]+BF4, used indifferently as precursors in palladium-catalyzed allylic substitutions, are not equivalent. In both situations, the mechanism of the Pd-catalyzed allylic substitution depends on the concentration of the chloride ions, delivered by the precursor or purposely added, that determines which species, [(η3-allyl)PdL2]+ or/and η1-allyl-PdClL2 are involved in the nucleophilic attack with consequences on the rate of the reaction and probably on its regioselectivity. Consequently, the chloride ions of the catalytic precursors [Pd(η3-allyl)(μ-Cl)]2 must not be considered as ‘innocent’ ligands.  相似文献   

15.
Published data on the structural characteristics of the local surrounding of lanthanide ions in aqueous solutions of respective salts obtained by different research methods under standard conditions are reviewed. Structural parameters like the coordination number, interparticle distance, parameters of the second coordination sphere, and types of ionic association are discussed.  相似文献   

16.
Ab initio-based charge partitioning of ionic systems results in ions with non-integer charges. This charge-transfer (CT) effect alters both short- and long-range interactions. Until recently, the effects of CT have been mostly neglected in molecular dynamics (MD) simulations. The method presented in this paper for including charge transfer between ions and water is consistent with ab initio charge partitioning and does not add significant time to the simulation. The ions of sodium, potassium, and chloride are parameterized to reproduce dimer properties and aqueous structures. The average charges of the ions from MD simulations (0.900, 0.919, and -0.775 for Na(+), K(+), and Cl(-), respectively) are consistent with quantum calculations. The hydration free energies calculated for these ions are in agreement with experimental estimates, which shows that the interactions are described accurately. The ions also have diffusion constants in good agreement with experiment. Inclusion of CT results in interesting properties for the waters in the first solvation shell of the ions. For all ions studied, the first shell waters acquire a partial negative charge, due to the difference between water-water and water-ion charge-transfer amounts. CT also reduces asymmetry in the solvation shell of the chloride anion, which could have important consequences for the behavior of chloride near the air-water interface.  相似文献   

17.
18.
The effects of concentration and an oriented external electric field on the transformations of hydrogenbonded structures of trimesic acid(TMA) and terephthalic acid(TPA) have been investigated at a liquidsolid interface by scanning tunneling microscopy(STM).The triangular periodic TMA framework can be transformed into a flower-like structure by changing the STM sample bias sign in situ.Networks of TMA and TPA are porous at a negative substrate bias,but typically change to relatively compact forms when the polarity of the applied bias is reversed.This change is reversible if the applied bias is reversed.The effects have potentials to locally control the capture and release of analytes in host-guest systems and the 2D morphology in multicomponent layers.  相似文献   

19.
We analyse transient effects in an electrolytic cell submitted to an external voltage and determine the relaxation time of the redistribution of the ions and of the potential. We consider the case in which adsorption effects at the interface with the electrodes are present and show that the typical relaxation time, for applied voltage V 025 mV, is of the order of tens of seconds for commercial nematic liquid crystals. When V 0>25 mV the linearized analysis is no longer valid. In this case, the relaxation time depends on the applied voltage. Increasing V 0, but still remaining in the low amplitude limit, the relaxation time starts increasing. This indicates that the reduction of the actual field in the sample, for moderate values of V 0, plays an important role. For large values of V 0, the relaxation time is a decreasing function of V 0. This result is interpreted in terms of a simple model, according to which the ionic charge is localized in a surface layer whose thickness depends on the amplitude of the applied voltage.  相似文献   

20.
The structure of the electric double layer of charged nanoparticles and colloids in monovalent salts is crucial to determine their thermodynamics, solubility, and polyion adsorption. In this work, we explore the double layer structure and the possibility of charge reversal in relation to the size of both counterions and coions. We examine systems with various size-ratios between counterions and coions (ion size asymmetries) as well as different total ion volume fractions. Using Monte Carlo simulations and integral equations of a primitive-model electric double layer, we determine the highest charge neutralization and electrostatic screening near the electrified surface. Specifically, for two binary monovalent electrolytes with the same counterion properties but differing only in the coion's size surrounding a charged nanoparticle, the one with largest coion size is found to have the largest charge neutralization and screening. That is, in size-asymmetric double layers with a given counterion's size the excluded volume of the coions dictates the adsorption of the ionic charge close to the colloidal surface for monovalent salts. Furthermore, we demonstrate that charge reversal can occur at low surface charge densities, given a large enough total ion concentration, for systems of monovalent salts in a wide range of ion size asymmetries. In addition, we find a non-monotonic behavior for the corresponding maximum charge reversal, as a function of the colloidal bare charge. We also find that the reversal effect disappears for binary salts with large-size counterions and small-size coions at high surface charge densities. Lastly, we observe a good agreement between results from both Monte Carlo simulations and the integral equation theory across different colloidal charge densities and 1:1-electrolytes with different ion sizes.  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