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.
Pseudopotential model was constructed to simulate the H3O+(H2O) n Cl clusters at room and stratosphere temperatures using the Monte Carlo method. Numerical values of interaction parameters were restored from the experimental data on the free energy and entropy of vapor nucleation on ions in the combination with the data of quantum chemical calculations of the optimal configurations of HCl(H2O) n clusters. The stability of various cluster structures and the probability of the rupture of intramolecular HCl bond in these clusters were analyzed.  相似文献   

2.
The free energy, entropy, and work of formation of H3O+(H2O)n clusters (n=1–27) in water vapor (300 K) were calculated by the Monte Carlo method. Binary correlation functions were calculated. The calculations are based on the nonpair interaction model presented in the previous publication. The hydration shell of the ion is thermally stable in the size range under study. Nonpair interactions exert an essential effect on the structure of the cluster. Fitting the cluster behavior to its experimental thermodynamic characteristics shows that the excess charge of the ion is spatially delocalized at room temperature, and the role of hydrogen bonds is strengthened on this background. Clusters formed on electric charges have such a fundamental characteristic as transition size. The transition size is independent of vapor pressure and demarcates two qualitatively different mechanisms of holding molecules in a cluster. A change in the holding mode is reflected on the mechanism of vapor nucleation.Translated from Zhurnal Obshchei Khimii, Vol. 74, No. 10, 2004, pp. 1585–1592.Original Russian Text Copyright © 2004 by Shevkunov.For communication I, see [1].This revised version was published online in April 2005 with a corrected cover date.  相似文献   

3.
Measurements are reported on the formation of negative ions in O2, O2/Ar and O2/Ne clusters aimed at establishing the mechanisms of anion formation and the role of inelastic electron scattering by the cluster constituents on negative ion formation in clusters. In the case of pure O2 clusters the main anions we detected are of two types: O(O2) n0 and (O2) n 1– . The yields of O(O2) n showed maxima at 6.3, 8.0 and 14.0 eV and the data suggest O as their precursor; the maxima at 8 and 14 eV are due to the production of O via symmetry forbidden dissociative attachment processes in O2 at these energies which become allowed in clusters. The yields of (O2) n showed a strong maximum at near-zero energy (0.5 eV) and also at 6.3, 8 and 14 eV. With the exception of the near-zero energy resonance, the (O2) n anions at 6.3, 8 and 14 eV are attributed to nondissociative attachment of near-zero energy secondary electrons to O2 clusters. The slow secondary electrons result predominantly from scattering via the O 2 negative ion states of incident electrons with energies in their respective regions. Similar results were obtained for the mixed O2/rare gas clusters except that now a feeble and distinctly structured contribution in the yields of O(O2) n , (O2) n (and Ar(O2) n ) was observed at energies >10 eV. These anions are believed to have the lowest negative ion states of Ar* (Ne*) as their precursors.  相似文献   

4.
The molecular structure of Cl(H2O) n clusters, n = 1–60, in equilibrium with vapor, and the cluster with n = 500 was studied by the Monte Carlo method. The first hydrated layer of a cluster is formed in unsaturated water vapors. The second hydrated layer begins to be formed in saturated vapor. The position of hydrated layers is not changed with an increase in cluster size and coincides with the position of the hydrated layers of ions in aqueous solutions of weak electrolytes. Orientational order in a cluster also has the layered structure. The orientation of molecules between the layers is random. The stability of the first layer is ensured only due to direct interactions with ions, whereas the stability of subsequent layers is due to cooperative interactions between molecules and between molecules and ions. As temperature decreases, the effect of ion displacement to the cluster surface becomes stronger.  相似文献   

5.
The addition of hydrogen in the reaction atmosphere is effective in promoting the activity of Ag/alumina and Ag-zeolites on the selective reduction of NO by hydrocarbons (HC-SCR) at low temperatures. The increment of NO conversion over Ag-MFI corresponds to the periodic addition of hydrogen into C3H8-SCR conditions. The UV–VIS spectra of Ag-MFI have revealed that the addition of hydrogen results in the formation of Agnδ+ clusters due to partial reduction and agglomeration of Ag species. The coincidence of the formation of the Agnδ+ clusters and the increment of NO conversion suggests that Agnδ+ clusters are the highly active species for HC-SCR. From analysis by H2-TPR, UV–VIS, and EXAFS, the structure of Agnδ+ clusters on Ag-MFI is identified as being Ag42+ on average. The formation of Ag clusters was strongly affected by the type of zeolites: The major Ag species are Ag+ ions for MOR, Agnδ+ clusters for MFI and BEA, and relatively large metallic Agmparticles for Y. The sequence of Ag agglomeration (MOR < MFI < BEA < Y) is in accordance with the strength of the acid sites of zeolites. It can be expected that the interaction between the positive charge of Agnδ+ clusters and acid sites, i.e., the ion-exchange site of zeolites, stabilizes Agnδ+ clusters. The type of Ag species under HC-SCR conditions depends on the concentration of gas-phase oxidants (NO, O2) and reductants (H2, HC), and also on the number and strength of the zeolite acid sites.  相似文献   

6.
The structure of nearly saturated or supersaturated aqueous solutions of NaCI [6.18 mol (kg H2O)–1], KCI [4.56 mol (kg H2O)–1], KF [16.15 mol (kg H2O)–1] and CsF [31.96 mol (kg H2O)–1] has been investigated by means of solution X-ray diffraction at 25°C. In the NaCI and KCI solutions about 30% and 60%, respectively, of the ions form ion pairs and the Na+–Cl and K+–Cl distances have been determined to be 282 and 315 pm, respectively. The average hydration numbers of Na+ and Cl ions are 4.6 and 5.3, respectively, in the NaCI solution and those of K+ and Cl ions in the KCI solution are both 5.8. In the KF solution, clusters containing some cations and anions, besides 1:1 (K+–F) ion pairs, are formed. The K+–F interatomic distance has been determined to be 269 pm, and nonbonding K+...K+ and F...F distances in the clusters are 388 and 432 pm, respectively, and the average coordination numbers n KF , n KK and n FF have been estimated to be 2.3, 1.9, and 1.6, respectively. In the highly supersaturated CsF solution an appreciable amount of clusters containing several caesium and fluoride ions are formed. The Cs+–F distance in the cluster has been determined to be 312 pm, while the nonbonding Cs+...Cs+ and F...F distances are estimated to be 442 and 548 pm, respectively, the distances being about and times the Cs+–F distance, respectively. The coordination numbers n CsF , n CsCs , and n FF in the first coordination sphere of each ion are 3.3, 2.3 and 5.3, respectively, and the result shows the formation of clusters of higher order than 1:1 and 2:2 ion pairs. These ion pairs and clusters may be regarded as embryos for the formation of nuclei of crystals and the results obtained in the present diffraction study support observations for the nucleation of the alkali halide crystals studied by molecular dynamics simulations previously examined.  相似文献   

7.
The energy of coordination-induced stabilization and the enthalpy of formation of gaseous metal closo-heteroclusters of the M@N k B r C s n type (m = k + r + s = 12, 24, or 28; n = 0–4), where M = Li, Mg, Al, Ti, Zr, Hf, V, Nb, Mo, Ru, Rh, Ir, Ta, Pt, Pd, and Au, were estimated in terms of a structural-thermodynamic model. The stabilizing role of metals in clusters was demonstrated. The energies D 0 of M–N, M–B, and M–C bonds were found to be underestimated by the MO LCAO method at the HF/6-31G* level.  相似文献   

8.
The characterization of the clusters formed on alkaline hydrolysis of [PdCl4]2– was performed using17O,23Na,35Cl,133Cs NMR and UV spectroscopy. The chemical composition of the clusters was found to be [Pd(OH)2] n ·nNaCl. No mononuclear oxo- or hydroxocomplexes were detected. The spatial structure of the clusters is stabilized by alkali metal cations.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 4, pp. 675–679, April, 1993.  相似文献   

9.
CnS (1 ≤ n ≤ 20) clusters have been investigated by means of the density functional theory. As a general rule, when 1 ≤ n ≤ 17 the energetically most favorable isomers are found to be the linear arrangement of nuclei (Cv) with the sulphur atom at the very end of the carbon chain. The electronic ground state is alternately predicted to be 1+ for odd n or 3 for even n with a conspicuous odd–even effect in the stability of these clusters. The C18S cluster is predicted to have a S-capped monocyclic structure (1A1), but with a low barrier to linearity. On the other hand, C19S and C20S are unambiguously linear in the 1+ and 3 electronic ground states, respectively.  相似文献   

10.
Lithium-excess binary clusters LinFn−1 (n=2–9) were detected by photoionization time-of-flight mass spectrometry in a supersonic cluster beam generated by a laser ablation of a solid mixture of lithium fluoride and nitride. Laser power dependence of the Li2F+ signal intensity has indicated that the ionization energy of the hyperlithiated Li2F molecule is lower than 4.66 eV. The theoretical vertical ionization energy obtained by the CCSD(T)/6-311+G(d)//B3LYP/6-311+G(d) calculations are 4.47 eV. No nitrogen-containing clusters were detected. The absence of Li4N is ascribed to the exothermicity of the reaction, 2Li3N→N2+Li6.  相似文献   

11.
The potential dependence of the differential capacitance is measured by an ac bridge at 420 Hz and 32°C at Pb–Ga/H2O interface in 0.05 M Na2SO4 solutions containing n-butanol in different concentrations. Adsorption parameters for n-butanol are determined using a regression analysis and compared with those for Hg, Tl–Ga, and Bi–Ga. As follows from this comparison, though there is no chemisorption interaction between these metals and water, the energy of n-butanol adsorption at these metals depends on the metal nature. The Pb–Ga data fit a common correlation dependence of the electronic capacitance of different electrodes (C m –1)phys on the energy of the n-butanol molecules adsorption thereon in the absence of a metal–water chemisorption interaction. This finding evidences that the dependence of the energy parameters for the adsorption of organic compounds on the metals' electronic properties, when characterized by (C m –1)phys, is of general nature.  相似文献   

12.
Electron attachment to clusters of acetone (A), trifluoroacetone (TFA) and hexafluoroacetone (HFA) is studied in a crossed beam experiment with mass spectrometric detection of the anionic products. We find that the electron attachment properties in A change dramatically on going from isolated molecules to clusters. While single acetone is a very weak electron scavenger (via a dissociative electron attachment (DEA) resonance near 8.5 eV), clusters of A capture electrons at very low energy (close to 0 eV). The final ionic products consist of an ensemble of molecules (M) subjected to the loss of two neutral H2 molecules ((Mn−2H2), n ≥ 2). Their formation at low energies can only be explained by invoking new cyclic structures and polymers. In clusters of TFA, anionic complexes containing non-decomposed molecules (Mn) including the monomer (M) and ionic products formed by the loss of one and two HF molecules are observed. Loss of HF units is also interpreted by the formation of new cyclic structures in the anionic system. HFA is a comparatively stronger electron scavenger forming a non-decomposed anion via a narrow resonant feature near 0 eV in the gas phase. In HFA clusters, the non-decomposed parent anion is additionally observed at higher electron energies in the range 3–9 eV. The M signal carries signatures of self-scavenging processes, i.e., inelastic scattering by one molecule and capture of the completely slowed down electron by a second molecule within the same cluster. The scavenging spectrum is hence an image of the electronically excited states of the neutral molecule.  相似文献   

13.
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   

14.
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   

15.
We develop and test an approximate approach for canonical simulations of weakly bound atomic or molecular systems for which some degrees of freedom can be treated separately by quantum mechanics. The system chosen for testing is Kr10–H, for which the adiabatic approximation applied to separate the hydrogen degrees of freedom works reasonably well. The hydrogen atom is bound to the Kr clusters at cold temperatures and we calculate several bound states for clusters in the n=1–9 range, in the global minimum configuration. The structural character of the mixed quantum classical simulation is substantially different than the classical simulation for Kr10–H as a result of zero point energy effects. When quantum effects are included, the low temperature dynamics of Kr10–H are dominated by a significant well to well hopping about an incomplete icosahedral krypton core.  相似文献   

16.
The equilibrium structures, binding energies, and vibrational spectra of the clusters CH3F(HF)1 n 3 and CH2F2(HF)1 n 3 have been investigated with the aid of large-scale ab initio calculations performed at the Møller–Plesset second-order level. In all complexes, a strong C–FH–F halogen–hydrogen bond is formed. For the cases n = 2 and n = 3, blue-shifting C–HF–H hydrogen bonds are formed additionally. Blue shifts are, however, encountered for all C–H stretching vibrations of the fluoromethanes in all complexes, whether they take part in a hydrogen bond or not, in particular also for n = 1. For the case n = 3, blue shifts of the ν(C–H) stretching vibrational modes larger than 50 cm−1 are predicted. As with the previously treated case of CHF3(HF)1 n 3 complexes (A. Karpfen, E. S. Kryachko, J. Phys. Chem. A 107 (2003) 9724), the typical blue-shifting properties are to a large degree determined by the presence of a strong C–FH–F halogen–hydrogen bond. Therefore, the term blue-shifted appears more appropriate for this class of complexes. Stretching the C–F bond of a fluoromethane by forming a halogen–hydrogen bond causes a shortening of all C–H bonds. The shortening of the C–H bonds is proportional to the stretching of the C–F bond.  相似文献   

17.
Förster–Dexter theory for resonant energy transfer is extended to higher order and applied to explain the rates of energy transfer and migration processes in highly forbidden transitions for some solid-state lanthanide (Ln) ion systems for which experimental results are available. The second-order two-body energy transfer mechanism involves two inter-ion correlated dipole electrostatic interactions, i.e. dipole dipole–dipole dipole (dd–dd) energy transfer, also termed Axe–Axe energy transfer in view of the similarity of the theoretical formalism with that for two-photon transitions. Each of the dipolar transitions consists of a transition from the 4fn configuration to an opposite-parity configuration, taken to be 4fn−15d. dd–dd energy transfer is a short-range (R−12) interaction so that it is most important in systems with short donor Ln–acceptor Ln separations. The energy transfer formalism is extended to include spin-forbidden transitions at one or two sites, the so-called Axe–Judd–Pooler (Axe–JP) and JP–JP energy transfer. In some cases the dd–dd mechanism is the dominant energy transfer process, as exemplified herein for energy migration in the 5D0 state of Sm2+ in SrF2, and also in the 5D0 state of Eu3+ in Cs2NaEuCl6.  相似文献   

18.
Modified Pencil Lead as a New Fiber for Solid-Phase Microextraction   总被引:2,自引:0,他引:2  
Dj. Djozan  Y. Assadi 《Chromatographia》2004,60(5-6):313-317
The efficiency of modified pencil lead as a new fiber for solid-phase microextraction (SPME) has been investigated. Modified pencil lead fibers have been prepared by use of several activation processes, for example heating at 600 °C in the stream of inert gas (He), heating under reflux with concentrated H2SO4 , fusing with NaOH at 400 °C, and activation at 600 °C with water vapor for 60 min. The fibers were used for extraction of trace amounts of polycyclic aromatic hydrocarbons (PAH) from aqueous samples. Monitoring of extracted compounds and quantitative analysis of model samples were performed by capillary GC–FID. The results obtained prove the suitability of the proposed fibers for sampling organic compounds from water. Effects on extraction efficiency of factors such as temperature, salting out, stirring speed, and exposure time were studied. Under optimum conditions and using one fiber for extraction of naphthalene as a typical compound, a relative standard deviation of 5.3% (n=7) was achieved. The calibration plot was linear in the range 50–10,000 pg mL–1 (r=0.9997) and the detection limit was 25 pg mL–1 (S/N=3). This fiber is very stable at high temperature, inexpensive, and can be prepared simply.  相似文献   

19.
Photochemical transformations of phenothiazine (PTA) in solutions of halomethanes CHnX4–n (X = Cl, Br; n = 0, 1, 2) and in n-hexane—CHnX4–n mixtures under the irradiation with = 337 and 365 nm were studied. The rate constants of quenching of PTA fluorescence with halomethanes (k q) are 4·105—1.3·1010 L mol–1 s–1. The process occurs due to electron transfer with the C—X bond cleavage in the radical anion fragment of the primary radical ion pair. This results in the formation of the stable radical cation salt (PTA·+X). The plot of k q vs. free energy of electron transfer corresponds to the Rehm—Weller empirical equation for a one-electron process and is satisfactorily described in terms of the theory of nonradiative electron transitions in the approximation of one quantum vibration.  相似文献   

20.
Layered crystalline zirconium phenylphosphonate, Zr(O3PC6H5)2, changed its interlamellar distance of 1481 pm after intercalation of n-alkylmonoamines, CH3---(CH2)n---NH2 (n=0–6). The infrared spectra of the precursor host and the corresponding intercalated compounds presented vibrations associated with PO3 groups in the 1163–1039 cm−1 range and additional bands related to C---H stretching bands in the 2950–2850 cm−1 interval were observed after amine insertion. The thermogravimetric curves showed a mass loss assigned to the phenyl group; however, the amine intercalated fraction was not quantitatively determined. A peak in the 31P NMR spectrum centered at −6 ppm for the host was observed. The surface area was 42.0±0.2 m2 g−1 and the scanning electron micrograph gave images consistent with lamellar structural features. The layered compound was calorimetrically titrated with amine in ethanol, requiring three independent operations: (i) titration of matrix with amine, (ii) matrix salvation, and (iii) dilution of the amine solution. From those thermal effects the variation in enthalpy was calculated as: −41±1.00,−33.28±0.50,−34.40±0.80,−10.40±0.40,−12.40±0.42,−16.10±0.08 and −7.0±0.04 kJ mol−1, for n=0–6, respectively. The exothermic enthalpic values reflected a favorable energetic process of amine–host intercalation in ethanol. The negative Gibbs free energy results supported the spontaneity of all these intercalation reactions. The positive favorable entropic values, as carbon chain size increased, are in agreement with the free solvent molecules in the medium, as the amines are progressively bonded to the crystalline lamellar inorganic matrix at the solid/liquid interface.  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