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 957 毫秒
1.
The electronic and geometric structures, energy stabilities, normal mode frequencies, and spin density distributions (in radicals) of different stepwise-chlorinated aluminum clusters Al13Cl n ? (n = 1–9) are calculated within the B3LYP approximation of the density functional theory using 6-31G* and 6-311+G* basis sets. The results are compared with analogous computation data on hydrides Al13H n ? (n = 1–12) obtained at the same level. The general qualitative pattern for related series of hydrides, chlorides, and iodides (as well as fluorides and bromides) turns out to be similar in many respects. For all Al13X n ? clusters with different electronegative substituents X, there is a set of a considerable number of low-lying closely spaced inner isomers (with a centered icosahedral cage), marquee isomers, and outer isomers (capped). The effects found by calculations in centered icosahedral isomers—localization of spin density on the trans-Al* atom in radical anions and its associated trans addition rule for an even substituent and the zigzag (odd-even) dependence of the energies D n (X) of successive addition of substituents X to the metal cage on n described in the framework of the molecular model of the valence states of the Al 13 ? superatom—should also be shared by many Al13X n ? series with different X’s. The differences between hydrides Al13H n ? and chlorides Al13Cl n ? of the same type are quantitative. For the hydrides, inner isomers are preferable in the first half of the series (n = 1–6); and in the second half (n = 7–12), outer isomers are more favorable. For the chlorides, icosahedral isomers are preferable only at the very beginning of the series. In the other cases, nonicosahedral structures are most favorable, for which the situation becomes very complicated due to the large number of position isomers and the aforementioned simple rules found for centered icosahedral structures are fulfilled to a considerably less extent or not at all.  相似文献   

2.
The vaporization of the NaI-PrI3 quasi-binary system was studied by high-temperature mass spectrometry over the whole concentration range. At 623–994 K, saturated vapor contained not only (NaI) n and (PrI3) n molecules (n = 1, 2) and Na+(NaI) n (n = 0–4) and I?(PrI3) n (n = 1–2) ions but also mixed molecular and ionic associates recorded for the first time (NaPrI4, Na2PrI5, NaPrI 3 + , Na2PrI 4 + , Na3PrI 5 + , Na4PrI 6 + , NaPrI 5 ? , and NaPr2I 8 ? ). The partial vapor pressures of molecules were calculated, and the equilibrium constants of the dissociation of neutral and charged associates were measured. The enthalpies of molecular and ion-molecular reactions were determined, and the enthalpies of formation of gaseous molecules and ions were obtained.  相似文献   

3.
The energies and structural and spectroscopic characteristics of endohedral (MO4©B20O 30 n? ) and exohedral (MO4 · B20O 30 n? ) isomers of oxoborate complexes with MO 4 n? tetraoxo anions with 32 valence electrons located in the inner and outer spheres of the B20O30 cluster have been calculated by the density functional theory method (B3LYP). It has been demonstrated that, among the endohedral MO4©B20O 30 n? clusters with strong multiply charged anions (VO 4 3? , CrO 4 2? , PO 4 3? , SO 4 2? , AsO 4 3? , SeO 4 2? , etc.), the isomer in which a “guest” tetrahedron MO4 is located at the center of the B20O30 cage and bonded to it through internal oxygen bridges M-O*-B is the most favorable one. Among the exohedral analogues MO4 · B20O 30 n? , two most favorable isomers contain the “capping” MO4 tetrahedron bonded to the B20O30 cage through two and three external M-O-B bridges. For the complexes with doubly charged SO 4 2? and SeO 4 2? anions, the third exohedral isomer in which the sulfite or selenite group MO3 is bidentately coordinated to the oxidized B20O29(OO) cage with one peroxide bridge turns out to be close in energy to the above two isomers. For the systems with high negative charge n, the exohedral isomers are much more favorable than the endohedral isomer; however, with decreasing charge, the difference in energy between them decreases to ~10–18 kcal/mol, so that the exo–endo transition between them can require moderate energy inputs. For the endohedral complexes with singly charged ClO 4 ? and BrO 4 ? anions, two isomers with close energies are preferable in which the central atoms of the guest tetrahedra are reduced to the state of singly charged ions, while the oxoborate cage is oxidized to B20O26(OO)4 with four peroxide groups B-O-O-B and retains its closed (closo) structure. In the most favorable isomer of the complexes with multicharged ortho-anions BO 4 5? , CO 4 4? , and NO 4 3? , the outersphere anion is reduced to, respectively, borate, carbonate, and nitrate bidentately coordinated to the oxidized B20O29(O)2 cage with an open structure and two strongly elongated terminal B-O bonds. The results are compared with the data of previous calculations of endohedral and exohedral vanadate complexes MO4©V20O 50 n? and MO4 · V20O 50 n? with the same guest anions MO 4 n? .  相似文献   

4.
5.
A DFT method with the B3LYP functional and the 6-311++G(d,p) diffuse basis set is used to predict geometries, relative stabilities, electronic structures, and the bonding of closo- and nido-GamBnmH n 2? , GemBnmH n m?2 , and AsmBnmH n 2 m?2 (n = 10, 12 and m = 1, 2) Clusters are obtained by replacing BH with isolobal GaH, GeH+, and AsH2+ fragments, keeping the same skeleton electron pairs (SEP). Based on the polyhedral skeletal electron pairs theory (PSEPT), closo and nido structures are predicted and can be of significant interest for experimentalists working in the field of heteroboranes. Different cluster stabilities are studied according to Gimarc′s and Williams′ rules, where our calculations show that the monosubstituted clusters deviate from these rules, giving rise to open structures. As2B8H n 2+ as 10-vertex structures lead to nido-type clusters, however, GemBnmH n m?2 (n = 10, 12 and m = 1, 2) give rise to closo isomers with close energies. All optimized structures exhibit large HOMO–LUMO gaps suggesting a good kinetic stability, thus predicting their isolation and characterization.  相似文献   

6.
The speed of sound (u), density (ρ), and viscosity (η) of 2,4-dihydroxyacetophenone isonicotinoylhydrazone (DHAIH) have been measured in N,N-dimethyl formamide and dimethyl sulfoxide at equidistance temperatures 298.15, 303.15, 308.15, and 313.15 K. These data were used to calculate some important ultrasonic and thermodynamic parameters such as apparent molar volume (V ? s st ), apparent molar compressibility (K ?), partial molar volume (V ? 0 ) and partial molar compressibility (K ? 0 ), were estimated by using the values of (V ? 0 ) and (K ?), at infinite dilution. Partial molar expansion at infinite dilution, (? E 0 ) has also been calculated from temperature dependence of partial molar volume V ? 0 . The viscosity data have been analyzed using the Jones–Dole equation, and the viscosity, B coefficients are calculated. The activation free energy has been calculated from B coefficients and partial molar volume data. The results have been discussed in the term of solute–solvent interaction occurring in solutions and it was found that DHAIH acts as a structure maker in present systems.  相似文献   

7.
The molecular and ionic sublimation of polycrystals and single crystals under Knudsen effusion and Langmuir evaporation conditions is reported. In both sublimation regimes, the sublimation product at 780–1050 K contains neodymium tribromide monomer and dimer molecules, as well as the negative ions NdBr 4 ? , Nd2Br 7 ? , and Br?. The dimer-to-monomer flux ratio j(Nd2Br6)/j(NdBr3)is larger in the molecular beam coming out of the effusion hole, while the ratio of the sublimation fluxes of the negative ions, j(Nd2Br 7 ? )/j(NdBr 4 ? ), is independent of the sublimation conditions. The partial pressures of the neutral components of the vapor have been determined, and the enthalpies and activation energies of sublimation of neodymium tribromide as monomer and dimer molecules and NdBr 4 ? and Nd2Br 7 ? ions have been calculated. The equilibrium constants of ion-molecule reactions have been measured, and the enthalpies of these reactions have been determined. Based on these data, values of the thermodynamic properties Δ s H 0(298.15) and Δ f H 0(298.15) are recommended for the monomer and dimer molecules and the NdBr 4 ? and Nd2Br 7 ? ions.  相似文献   

8.
The substitution equilibria AuCl 2 ? + iNH 4 + = Au(NH3)iCl2 ? i + iCl? + iH+, β i * . were studied pH-metrically at 25°C and I = 1 mol/L (NaCl) in aqueous solution. It was found that logβ 1 * = ?5.10±0.15 and logβ 2 * = ?10.25±0.10. For equilibrium AuNH3Clsolid = AuNH3Cl, log K s = ?3.1±0.3. Taking into account the protonation constants of ammonia (log K H = 9.40), the obtained results show that for equilibria AuCl 2 ? + iNH3 = Au(NH3)iCl2 ? i + iCl?, logβ1 = 4.3±0.2, and logβ2 = 8.55±0.15. The standard potentials E 0 1/0 of AuNH3Cl0 and Au(NH3) 2 + species are equal to 0.90±0.02 and 0.64±0.01 V, respectively.  相似文献   

9.
Substitution of chloride ions in AuCl 4 ? with ethylenediamine (en) and propylenediamine (tn) is studied by capillary zone electrophoresis at I = 0.05 M and T = 25°C. The substitution constants are determined: AuenCl 2 + + en = Auen 2 3+ + 2Cl, logK2 = 10.4; AuCl 4 ? + tn = AutnCl 2 + + 2Cl, logK1 = 16.1; AutnCl 2 + + tn = Autn3+2 + 2Cl, logK2 = 12.0.  相似文献   

10.
The lowest energy structures and electronic properties of ErSi n (n = 3–10) and their anions were probed using the ABCluster global search technique combined with the PBE, TPSSh and B3LYP schemes. The lowest energy energies of neutral ErSi n (n = 3–10) can be regarded as substituting a Si atom of the lowest energy structure of Sin+1 with a Er atom. The additional electron effects on the geometries are very strong, resulting the lowest energy structures of ErSi n ? with n > 6 are different from their neutral counterparts. Starting from n = 7, the potential energy surfaces of ErSi n ? are very flat, resulting isomeric arrangements occur and functional dependence of the predicted most stable structures exist. The AEAs, VDEs and simulated PES of ErSi n (n = 3–10) are reported. Introducing Er to Si cluster can significantly improve photochemical reactivity of the cluster. The 4f electron of Er atom in ErSi4, ErSi n ? (n = 4, 7–10) prefers to take part in bonding. The total magnetic moments of ErSi n and their anions are mainly provided by the 4f electrons of Er atom. The dissociation energies of Er from ErSi n and their anions were evaluated to inspect relative stability.  相似文献   

11.
Density functional theory was used to study model ethylene reactions with CpTiIIIEt+A? (A? = CH3B(C6F5) 3 ? , or B(C6F5) 4 ? ; A? can be absent) compounds. The polymerization of ethylene on an isolated CpTiEt+ cation is hindered because of equilibrium between the CpTi(C2H4)Et+ primary complex and the primary product of CpTiBu+ insertion. At the same time, the polymerization of ethylene on CpTiEt+A? ion pairs (A? = CH3B(C6F5) 3 ? or B(C6F5) 4 ? ) is thermodynamically allowed (ΔE from ?26.2 to ?25.6 kcal/mol and ΔG 298 from ?10.9 to ?10.4 kcal/mol) and is not related to overcoming substantial energy barriers (ΔE # = 8.2?12.3 kcal/mol and ΔG 298 ) = 7.8?13.3 kcal/mol). The degree of polymerization can be low because of the effective occurrence of polymer chain termination by hydrogen transfer from the polymer chain to the monomer.  相似文献   

12.
The mean atomic Gibbs energies of formation of (Δ f ? at 0 ) of s-, p-, and d-element diphosphates have been calculated using ion increments of the Gibbs energy (Δ f G 0). The diphosphate hydrolysis kinetics is considered, and a correlation between the Δ f ? at 0 values and the hydrolysis rate constants is presented.  相似文献   

13.
Polarograms for the reduction of glycinate complexes of palladium(II) (5 × 10?5 M) are obtained in equilibrium solutions of pH 0.8–3.0 with different protonated-glycine concentrations c Hgly (supporting electrolyte, 0.5 M NaClO4). It is established that the irreversible wave of reduction of complexes Pd(gly)2 corresponds to the diffusion limiting current I d (2) . A similar wave at pH 1.5 and c Hgly = 0.005 M, as well as at pH 1.0 and c Hgly = 0.05–0.5 M is preceded by the diffusion limiting current I d (1) . Values of the I d (2) /I d (1) ratio are close to the ratio between equilibrium concentrations of Pd(gly)2] and [Pdgly+], calculated using the step stability constant for Pd(gly)2. This fact testifies to the reduction of complexes Pdgly+ in the vicinity of I d (1) and complexes Pd(gly)2, in the vicinity of I d (2) . At pH 0.8–1.2 and [H2gly+] = 1 × 10?4 to 5 × 10?3 there is observed the diffusion-kinetic limiting current of the first wave I 1 (1) , which increases with increasing [H+] and decreasing [H2gly+]. The nature of the slow preceding chemical stage that occurs during the reduction of complexes Pdgly+ is discussed.  相似文献   

14.
The hydrolysis kinetics of the anion in 3d-element cyclotetraphosphates is considered. The thermodynamic functions of formation (Δ f H 0, Δ f G 0, and Δ f ? at 0 ) of the cyclotetraphosphates are calculated using the ion increment method. A linear correlation is established between and log K Δ f ? at 0 for these compounds.  相似文献   

15.
The total limiting molar electrical conductivities of ions and triads of ions and the association constants of ions with the formation of ion pairs and triads of ions were calculated from the concentration dependences of the electrical conductivity of solutions of lithium and sodium perchlorates in tetrahydrofuran at 278.15–318.15 K with the use of the method specially developed earlier. The experimental total limiting electrical conductivities were used to calculate the limiting molar electrical conductivities and attraction friction factors of separate ions (Li+, Na+, ClO 4 ? , Li2ClO 4 + , Na2ClO 4 + , Li(ClO4) 2 ? , and Na(ClO4) 2 ? ). The constants of ion association into ion pairs were used to calculate the Gibbs energy of non-Coulomb interionic interaction (ΔG*+?), and the constants of association into triads of ions, to determine the a 3 distance parameter between the centers of the ion and the dipole of the ion pair. Positive ΔG*+?), values and deviations of the experimental a 3 value from the distance parameter calculated theoretically (a 3 0 ) for the triad of ions (Δa 3 = a 3 ? a 3 0 ) were related to non-Coulomb repulsion in the region of overlap of the solvation shells of ions and the influence of temperature and ion charge density on this repulsion.  相似文献   

16.
A method has been purposed to calculate some of the thermodynamic quantities for the thermal deformation of a smectite without using any basic thermodynamic data. The Hanç?l? (Keskin, Ankara, Turkey) bentonite containing a smectite of 88% by volume was taken as material. Thermogravimetric (TG) and differential thermal analysis (DTA) curves of the sample were obtained. Bentonite samples were heated at various temperatures between 25–900°C for the sufficient time (2 h) until to establish the thermal deformation equilibrium.Cation-exchange capacity (CEC) of heated samples was determined by using the methylene blue standard method. The CEC was used as a variable of the equilibrium. An arbitrary equilibrium constant (K a) was defined similar to chemical equilibrium constant and calculated for each temperature by using the corresponding CEC-value. The arbitrary changes in Gibbs energy (ΔG a 0 ) were calculated from K a-values. The real change in enthalpy (ΔH 0) and entropy (ΔS 0) was calculated from the slopes of the lnK vs. 1/T and ΔG vs. T plots, respectively. The real changes in Gibbs energy (ΔG 0) and real equilibrium constant (K) were calculated by using the ΔH 0 and ΔS 0 values. The results at the two different temperature intervals are summarized as below: ΔG 1 0 H 1 0 S 1 0 T=?RTlnK 1=47000?53t, (200–450°C), and ΔG 2 0 H 2 0 S 2 0 T=?RTlnK 2=132000?164T, (500–800°C).  相似文献   

17.
The densities, viscosities and refractive indices of N,N /-ethylene-bis(salicylideneiminato)-diaquochromium(III) chloride, [Cr(salen)(H2O)2]Cl, in aqueous dimethylsulfoxide (DMSO) with different mass fractions (w 2 = 0.20, 0.40, 0.60, 0.80 and 1.00) of DMSO were determined at 298.15, 308.15 and 318.15 K under atmospheric pressure. From measured densities, viscosities and refractive indices the apparent molar volumes (V φ ), standard partial molar volume (V φ 0 ), the slope (S V * ), standard isobaric partial molar expansibility (φ E 0 ) and its temperature dependence (?φ E 0 /?T) p , the viscosity B-coefficient, its temperature dependence (?B/?T), solvation number (S n ) and apparent molar refractivity (R D φ ), etc., were calculated and discussed on the basis of ion–ion and ion–solvent interactions. These results revealed that the solutions are characterized by ion–solvent interactions rather than by ion–ion interactions and the complex behaves as a long range structure maker. Thermodynamics of viscous flow was discussed in terms of transition state theory.  相似文献   

18.
The stepwise substitution equilibrium AuCl 2 ? +iX?=AuCl2?i X i ? +iCl?, βi, where X? is the glycinate ion (H2N-CH2-COO?), i = 1 or 2, at 25°C in an aqueous solution with I = 1.0 mol/L (NaCl) has been studied pH-metrically. The corresponding constants are logβ1 = 3.60 ± 0.10, and logβ2 = 6.2 ± 0.2.  相似文献   

19.
A convenient method is suggested for calculating thermally averaged powers of the normal vibrational coordinates Q i by iteratively solving the Bloch integral equation with an anharmonic function of potential energy using multidimensional Hermite polynomials. Analytical formulas of the first approximation regarding anharmonicity constant have been obtained for the following moments of thermally averaged density: 〈Q 1〉, 〈 Q 1 2 〉, 〈Q 1 Q 2〉, 〈Q 1 3 〉 〈Q 1 3 〉, 〈Q 1 Q 2 Q 3〉, 〈Q 1 4 〉, 〈Q 1 2 Q 2 2 〉, 〈Q 1 Q 2/3〉, 〈Q 1 Q 2 Q 3 2 〉, 〈 Q 1 Q 2 Q 3 Q 4〉.  相似文献   

20.
Dimethylgold(III) complexes with 8-hydroxyquinoline Me2Au(Ox) (I) and 8-mercaptoquinoline Me2Au(Tox) (II) were synthesized and studied. Complex II obtained for the first time was identified from the elemental analysis, IR, 1H NMR, and mass spectrometry data. The thermal properties of complexes I, II in condensed state were investigated by thermography. The temperature dependences of the saturated vapor pressure over crystals were measured by the Knudsen effusion method with mass spectrometric recording of the gas phase composition and the thermodynamic characteristics of the sublimation process were determined: for I, log P[Torr] = (14.6 ± 0.3) ? (6.34 ± 0.10) × 103/(T, K), Δ H subl o = 121.2 ± 1.9 kJ?1, Δ S subl o = 224.1 ± 4.6 J mol?1 K?1 (the temperature interval under study 80–115°C); for II, log P [Torr] = (13.3 ± 0.2) ? (6.30 ± 0.09) × 103/(T, K), Δ H subl o = 120.5 ± 1.7 kJmol?1, ΔS subl o = 199.3 ± 3.0 J mol?1 K?1 (86–145°C).  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