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.
Molecular dynamics simulations are performed to study oxygen adsorption on platinum clusters supported on a graphite surface. The Sutton–Chen many-body potential is used for the Pt–Pt interaction, whereas a Steele potential was used to represent the carbon surface. The oxygen–oxygen intramolecular force is modeled by a harmonic oscillator model and other interactions are described by the Lennard–Jones potential. The results indicate an optimum loading of platinum for maximum specific adsorption of oxygen. Adsorption isotherms are constructed and the energies and orientation of adsorbed oxygen are reported. The relevance of this study to electrode processes is discussed.  相似文献   

2.
Starting from the Hohenberg–Kohn theorem, atomic and molecular energies have been expressed rigorously as functionals of the electronic electrostatic potential, Velec(r). Explicit formulations have been derived for the functionals representing the kinetic energy and electronic interaction contributions to the total energies.Acknowledgements. The assistance of Dr. Jane S. Murray is greatly appreciated.Contribution to the Jacopo Tomasi Honorary Issue  相似文献   

3.
The adsorption of bovine serum albumin (BSA) on fused silica at pH 4.7 was studied at the single molecules level by total-internal-reflection fluorescence microscopy. This pH value was the isoelectric point of BSA. At low [BSA] of 20 pM, protein molecules adsorbed as monomers. At intermediate [BSA] of 500 pM, protein molecules adsorbed as clusters of about five monomers on average. Both monomers and clusters had adsorption rate coefficients of the order 10−7 m s−1 and desorption rate coefficients of about 2 × 10−2 s−1. The respective steady-state coverage was about 10× higher than that at neutral pH, presumably because of the more favorable BSA–silica electrostatics. At pH 4.7 and with [BSA] higher than 100 nM, adsorption begot further adsorption to produce nonlinear isotherms. The coverage at 1 μM BSA was 2.5× that of the linearly extrapolated coverage. This suggests that at pH 4.7, solute–adsorbate affinity was the dominant factor that explains the enhanced adsorption observed in ensemble measurements.  相似文献   

4.
Various effective components of the intermolecular interaction of water containing aggregates are examined and their modeling, in terms of the fundamental physical properties of the involved partners, is discussed. We focus, in particular, on the evolution of these components in going from the simplest neutral rare gas–water aggregates to bulk water and ionic water solutions. The analysis singled out that the model chosen to represent the van der Waals interaction as the composition of the action of three dispersion/induction–attraction centres and found to be appropriate to describe the lighter He–H2O and Ne–H2O systems, is not adequate to describe the heavier Ar–H2O aggregate. It was found, instead, that by increasing the mass of the rare gas, other short range contributions to the interaction come into play. Moreover, it was also found that the water molecule tends to behave as a single centre as the strength of the interaction increases. This led to the development of an effective model potential suitable to describe water clusters in the range going from gaseous to condensed phase. The role of electrostatic contributions is also evaluated. The proposed potential model is tested by comparing molecular beam scattering and neutron diffraction experiments with results of molecular dynamics (MD) calculations.  相似文献   

5.
A total of 16 hydrogen-bonded complexes between the lowest energy tautomers of L-leucine and each base of RNA have been characterized at the DFT level of theory. The most stable complexes are formed by L-leucine and guanine. Considering backbone of RNA, the affinity order between the L-leucine and bases is guanine> cytosine> adenine≈uracil in perfect accordance with the experimentation. The interplay between the transformed H–X bonds’ structural parameters from two monomers to the dimer accompanying with the shifts of the frequency for the H–X stretching mode and interaction energies has been discussed.  相似文献   

6.
The potential energy surfaces for the interatomic interaction in the Li+HCOO system have been investigated byab initio methods within the rigid-molecule approximation. Analytical potential expressions were fitted to 133 calculated SCF energies for the Li+-HCOO interaction, 42 SCF energies for the Li+-Li+ interaction, and 332 SCF energies for the HCOO-HCOO interaction. The global minimum on the Li+-HCOO surface is –170 kcal/mol and corresponds to the lithium ion lying on the C2 axis of the formate ion at 2.2 Å from the carbon atom on the oxygen side. The cation-cation and anion-anion interactions are repulsive everywhere, although the potential surface is markedly anisotropic for the HCOO-HCOO interaction.  相似文献   

7.
Summary The application of plasma-based SNMS to salt samples was examined to reveal the analytical features of this technique for the analysis of environmental material, which is frequently composed of ionic compounds. 1016 s–1 cm–2 argon ions of 300–500 eV energy were used to sputter halides, sulfates, nitrates and carbonates. The mass spectra are dominated by atomic signals of all elements of the salt and contain additional minor signals of binary homo- and heteroelemental clusters. The latter, such as salt monomers of alkali halides, are useful for compound identification. In sputter equilibrium the atomic mass signals can be used for elemental quantitation with a matrix dependence of about ±30% for the detection factors of most elements. As the average elemental detection factors are shown to be governed by atomic ionization probabilities the erosion flux from salts comprises mainly atoms. Results indicate that thermal release of atoms as well as emission of clusters are the main cause of the matrix dependence. The relative yield of the clusters was found to be strongly enhanced with increasing mass difference of the combined ions. The sputter yield of NaCl was determined to be 3.3 atoms/projectile at 490 eV argon ion energy which results in a depth propagation of 5 nm/s for the given sputter conditions.  相似文献   

8.
Orientational ordering in the surface layer of water–methanol mixture clusters was investigated by the molecular dynamics method. The dependence of the local electric potential on the composition of the mixture was considered. The surface potential of water–methanol binary solution was estimated.  相似文献   

9.
Vibronic reduction is considered for Coulomb interaction between centers for Heisenberg and double exchange in multielectron mixed-valency d2-d3 clusters. A simplified multimode vibration model indicates how effects arise via vibrations from the correlation interaction between electron shells. Vibronic reduction in double exchange can alter the magnetic parameters for the cluster's ground state.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 28, No. 2, pp. 114–121, March–April, 1992.  相似文献   

10.
In their study of solvent-averaged ion-ion pair potentials, Pettitt and Rossky show that, at least for the given underlying Born-Oppenheimer level model, there is a remarkable well in the potential of the solventaveraged Cl–Cl interaction in water; it has a minimum below-kT and near 3.6Å, the Pauling diameter for the Cl ion. As Pettitt and Rossky also showed, the osmotic coefficient of aqueus NaCl solution as a function of ionic strength (I) could be insensitive to this remarkable feature, even when the well implies that there is a substantial concentration of dimers at 1M or higher stoichiometric concentration. Here it is argued that the thermodynamics of mixtures of aqueous NaCl with, say, aqueous NaClO4, should be sensitive to the presence of the well in the Cl–Cl interaction. Accordingly we apply the HNC approximation to calculate the mixing coefficients that characterize the excess free energy functions of several model aqueous NaCl–NaClO4 mixed electrolyte solutions. For reference we use a model (V6) in which all six ion-ion pair potentials have the simple vanilla form: charged soft spheres supplemented by a dielectric image correction and an adjustable Gurney term that represents an interaction of the solvation structures about the ions. The six Gurney coefficients were adjusted to tune the model (I) close to the experimental for pure NaCl, pure NaClO4, and an equimolar mixture, all in the range 0.1I2M. We find that the V6 model gives mixing coefficients that agree closely with experiment. The calculations were also made for five models that incorporate, in varying degrees, the Pettitt-Rossky pair potentials. Comparison of the mixing coefficients for these models with the experimental data leads to the conclusion that the Cl, Cl well is unrealistic, although a shallower well in the same place might be acceptable.  相似文献   

11.
The vibronic aspect of the Jahn-Teller effect is considered for multielectron mixed-valency dimer clusters. A basic intermediate-coupling set is used in the numerical diagonalization of the vibronic Hamiltonian. The temperature dependence of the vibronic magnetic moment is derived for symmetrical and distorted mixed-valency dimer clusters. Strong vibronic interaction can suppress double exchange and lead to an antiferromagnetic ground state for the cluster. Distortions that eliminate the inversion center localize the surplus electron and suppress the ferromagnetic effect caused by double exchange.Chemical Institute, Moldovian Academy of Sciences. Moldovian State University. Translated from Zhurnal Strukturnoi Khimii, Vol. 34, No. 2, pp. 11–19, March–April, 1993.  相似文献   

12.
The photoionization efficiency (PIE) spectra and ionization potentials are reported for scandium cluster monoxides (Sc n O,n=5–36). As found for other transition metal clusters, strong dependence of ionization potential on cluster size is found for small clusters, with the ionization potentials of larger clusters decreasing relatively smoothly with increasing size. The IPs are 0.6–0.8 eV lower than that predicted by conducting spherical droplet model.Performed at Argonne National Laboratory  相似文献   

13.
In aqueous solution, the interaction between sodium dodecyl sulfate (SDS) and poly(ethylene glycol) (PEG) results in the formation of small aggregates or clusters of SDS attached to the PEG polymer chain. Selectivity coefficients for exchange of two monovalent (N-methyl-4-cyanopyridinium cation and Tl+) and two divalent (methylviologen cation and Cu2+) counterions at the surface of SDS–PEG clusters, determined employing photophysical techniques, are similar, but not identical, to those for exchange at the surface of SDS micelles in the absence of PEG. The principal factor affecting ion exchange selectivity in SDS–PEG clusters does not appear to be aggregate size or surface charge density but rather the presence of poly(oxyethylene) subunits at the aggregate surface.  相似文献   

14.
The interaction of the dyes Safranin-O (SO) and Orange II (OII) with aqueous colloidal Q-CdS clusters, which emit single fluorescence bands with maximum wavelengths at 481 nm (excitonic band) or 559 nm (trapped band), has been studied. This was carried out by monitoring both the photodegradation of the dye in the presence of the clusters and the quenching of the clusters fluorescence by the dyes. The photolysis experiments were carried out by excitation either at 520 nm (the wavelength at which the dyes, but not the clusters absorb light) or at 350 nm (the wavelength at which the clusters strongly absorb light, and the dyes have absorbance minima). At 520 nm, photodegradation of SO could be observed, which follows a first-order kinetics (for trapped-band clusters) and a second-order kinetics (for excitonic-band clusters). For the excitation wavelength of 350 nm, photodegradation of either of the dyes could not be observed. The Stern–Volmer plots for the quenching of the excitonic band-clusters fluorescence by SO show an upward curvature, pointing to the occurrence of more than one species acting as the fluorescence quencher, whereas the Stern–Volmer plots for the quenching of the trapped band-clusters fluorescence by SO are linear, indicating that only one species acts as a fluorescence quencher. Lambert–Beer type plots (absorbance vs. concentration) are linear for SO in water and in trapped band-clusters solutions, but a similar study of SO in excitonic band-clusters solution show the occurrence of a new band, which can be assigned to a ground-state dimer of the dye. The latter can be used to explain both the upward curvature of the Stern–Volmer plots and the second-order kinetics observed for SO photodegradation in the SO-excitonic band-clusters system. The Stern–Volmer plots for the quenching of both fluorescence bands by OII are linear.  相似文献   

15.
Uracil–(H2O)n (n = 1–7) clusters were systemically investigated by ab initio methods and the newly constructed ABEEMσπ/MM fluctuating charge model. Water molecules have been gradually placed in an average plane containing uracil. The geometries of 38 uracil–water complexes were obtained using B3LYP/6-311++G** level optimizations, and the energies were determined at the MP2/6-311++G** level with BSSE corrections. The ABEEMσπ/MM potential model gives reasonable properties of these clusters when comparing with the present ab initio data. For interaction energies, the root mean square deviation is 0.96 kcal/mol, and the linear coefficient reaches 0.997. Furthermore, the ABEEMσπ charges changed when H2O interacted with the uracil molecule, especially at the sites where the hydrogen bond form. These results show that the ABEEMσπ/MM model is fine giving the overall characteristic hydration properties of uracil–water systems in good agreement with the high-level ab initio calculations.  相似文献   

16.
Stabilization energy of the (H2) n clusters (n = 2–8) was calculated as a sum of the SCF interaction energy and the semiempirical interaction correlation energy estimated according to Sinanolu and Pamuk. Optimum successive attachment of hydrogen molecules leads to the formation of a gas-phase solvation shell consisting of seven hydrogen molecules. Basis set effect has been found to be important with all clusters under study. The non-additivity effect was investigated with the (H2)4 cluster. Vertical ionization potentials of the clusters considered are predicted to be 0.4–0.6 eV lower than the ionization potential of the parent H2 molecule.  相似文献   

17.
Ions of gold monomer and clusters emitted from a liquid metal ion source were mass-selected, and deposited on cleaved HOPG (highly oriented pyrolytic graphite) surfaces and on amorphous carbon thin films at room temperature with the impinging energy E i from 0 to 500 eV. The coverage of deposited ions were 1/100 and 1/1000 monolayers on HOPG surfaces and 1/3 monolayers on carbon films. Scanning tunneling microscopy of the HOPG surfaces deposited with low impinging energy (E i<50 eV) revealed that large clusters with diameters ranging from 2 to 5 nm and height of 1–2 layers were present instead of isolated monomers and original clusters. When E i was higher than 100 eV, HOPG surfaces were damaged and only bumpy surfaces were observed by STM. Transmission electron microscopy of Au+-deposited carbon films showed the formation of clusters with diameter 0.5–20 nm, depending on the E i and the time elapsed after deposition.  相似文献   

18.
Conclusions The crystal structure of Pt4(CO)5(PEt3)4 was determined and the IR spectra of this compound in the VCO region characteristic for such clusters were studied. Four models for the interaction of CO with metals upon adsorption were proposed on the basis of the geometrical parameters obtained for the coordination of CO in this and other clusters. Calculation of the frequencies and shape of the vibrations of all four models showed that we must independently evaluate the force constants of the interaction of the CO groups with the remaining part of the complex in order to develop reliable spectral indicators for the specific type of CO coordination.Translated from Zhurnal Strukturnoi Khimii, Vol. 26, No. 2, pp. 84–93, March–April, 1985.  相似文献   

19.
The axial distribution of the electrostatic potentials ofan electron-cyclotron resonance (ECR) processing plasma confined in a do magnetic mirror geometry was characterized. The potential profiles far argon and helium at 8.0x 10–4 and 4.0 × 10–4 Torr were measured using electron emissive probes. The experimental measurements were then compared with the predictions of a one-dimensional, electrostatic, particle-in-cell computer code which runs on a personal computer. The potential profiles as predicted by the code showed good agreement with the experimental measurements.  相似文献   

20.
In a series of molecular dynamics (MD) runs on (KI)108 clusters, the Born–Mayer–Huggins potential function is employed to study structural, energetic, and kinetic aspects of phase change and the homogeneous nucleation of KI clusters. Melting and freezing are reproducible when clusters are heated and cooled. The melted clusters are not spherical in shape no matter the starting cluster is cubic or spherical. Quenching a melted (KI)108 cluster from 960 K in a bath with temperature range 200–400 K for a time period of 80 ps both nucleation and crystallization are observed. Nucleation rates exceeding 1036 critical nuclei m−3 s−1 are determined at 200, 250, 300, 350, and 400 K. Results are interpreted in terms of the classical theory of nucleation of Turnbull and Fisher and of Buckle. Interfacial free energies of the liquid–solid phase derived from the nucleation rates are 7–10 mJ m−2. This quantity is 0.19 of the heat of transition per unit area from solid to liquid, or about two-thirds of the corresponding ratio which Turnbull proposed for freezing transition. The temperature dependence of σsl(T) of (KI)108 clusters can be expressed as σsl(T)∝T0.34.  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