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.
2.
Barrelene, H–C(CH=CH)3C–H, is an unsaturated polycyclic hydrocarbon containing three isolated double bonds in a non-planar arrangement. We have studied the transmission of field effects through the barrelene framework by analyzing the small structural changes occurring in the phenyl group of many Ph–C(CH=CH)3C–X molecules, where X is a variable substituent. Molecular geometries have been determined by quantum chemical calculations at the HF/6-31G* and B3LYP/6-311++G** levels of theory. Comparison with the results obtained for the corresponding saturated molecules, the bicyclo[2.2.2]octane derivatives Ph–C(CH2–CH2)3C–X, reveals a small, but significant, field-induced π-polarization of the barrelene cage, especially when the remote substituent is a charged group. Additional evidence of π-polarization is obtained by comparing the electric dipole moments of the two sets of uncharged molecules. The structural variation of the barrelene cage caused by the variable substituent in Ph–C(CH=CH)3C–X molecules has also been investigated. It is much larger than that of the phenyl group and depends primarily on the electronegativity of the substituent. Particularly pronounced is the concerted variation of the non-bonded distance between the bridgehead carbons of the cage, r(C···C) 1 BARR , and the average of the three C–C–C angles at the cage carbon bonded to the variable substituent, α 1 BARR . A scattergram of r(C···C) 1 BARR versus the corresponding parameter for bicyclo[2.2.2]octane derivatives, r(C···C) 1 BCO , shows that the variation of r(C···C) 1 BARR becomes gradually less pronounced than that of r(C···C) 1 BCO as the electronegativity of the substituent increases.  相似文献   

3.
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   

4.
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   

5.
The structural features of 38 mononuclear d 2-Re(V) octahedral monooxo complexes (I–XXXVIII) with oxygen atoms of bidentate-chelating (O, P) ligands (L n ) are considered. The atoms O(L n ) are mostly in trans positions to O(oxo) ligands. In three compounds of general formula [ReO(Lmono)(L n )2] (XXXVI–XXXVIII), the O atoms of two L n ligands occupy both trans and cis positions to oxo ligands. In one complex, namely, in [ReO(L n )(L tri 11 )], n = 3 (XXXV), the atom O(L3) is in the cis position to the oxo ligand; the trans position to O(oxo) is occupied by the atom O(L tri 11 ).  相似文献   

6.
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   

7.
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   

8.
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   

9.
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   

10.
The thermal sila-Pummerer rearrangement of diastereomeric 2,3,3-trimethyl-1,3-thiasilinane S-oxides was studied. Introduction of the methyl group in the 2 position of 3,3-trimethyl-3-thiasilinane S-oxide slows down the rearrangement. When heated in CCl4, the trans isomer (2-Meeq, SOeq) converts into the cis isomer (2-Meeq, SOax) which rapidly rearranges into 2,2,7-trimethyl-1,6,2-oxathiasilepane. On the contrary, the isomeric 2,4,4-trimethyl-1,4-thiasilinane S-oxide is thermally stable up to 160°C in DMSO. The inversion at the sulfur atom in 2,3,3-trimethyl-1,3-thiasilinane S-oxides and 2,4,4-trimethyl-1,4-thiasilinane S-oxides under the action of triethyloxonium tetrafluoroborate was studied. The trans isomer of 2,3,3-trimethyl-1,3-thiasilinane S-oxide (2-Meeq, SOeq) forms with Et3O+BF 4 ? a salt which decomposes in two ways. The first involves recovery of the starting sulfoxide due to Sn2 substitution at the carbon atom of the ethoxy group, and the second, convertion into the cis isomer (2-Meeq, SOax) which rearranges into 2,2,7-trimethyl-1,6,2-oxathiasilepane. Under the same conditions, the cis isomer of 2,3,3-trimethyl-1,3-thiasilinane S-oxide (2-Meeq, SOeq) decomposes to form siloxanes. trans-2,4,4-Trimethyl-4-thiasilinane S-oxide (2-Meeq, SOeq) under the action of Et3O+BF 4 ? convers into the cis isomer (2-Meeq, SOax). The B3LYP/6-311G(d,p) theoretical analysis showed that the thermal inversion at the sulfur atom in the compounds studied has a high energy barrier.  相似文献   

11.
By means of the formula \(e^{ - ((r^{XH} - r_0^{XH} )/b^{XHX} )^{5/3} } + e^{ - ((r^{YH} - r_0^{YH} )/b^{YHY} )^{5/3} } = 1\) that characterizes the correlation between the parameters of XH?Y linear fragments (r 0 XH , r 0 YH are the bond lengths in free molecules, b XHX, b YHY are the dimensional coefficients) r XX(r XH) and r XY(r YH) dependences are obtained. Given the length of a hydrogen bridge formed by O, N, and F atoms, they enable us to find the proton position in the bridge. The definition “a quasi-symmetric hydrogen bond,” based on the invariance of the r XX distance when the proton shifts by 0.1 Å, is established to be applicable to OHO, FHF, NHN, and ClHCl fragments. It is shown that the hydrogen bridge length remains almost constant (exceeds the minimum length by no more than 0.1 Å) if its bond orders are above 0.1. Here the displacement of the central proton can reach 0.2–0.3 Å.  相似文献   

12.
By applying a three-dimensional holographic vector of the atomic interaction field (3D-HoVAIF) to express the structure of three classical peptide drugs, quantitative structure activity relationship (QSAR) models are built by the multiple linear regression. The accuracy of the proposed model is illustrated using Q LOO 2 (cross-validation) and r 2 (test set validation). Moreover, the r m 2 metrics is used to further refine the predictive ability of the developed QSAR models. The results show that 3D-HoVAIF, due to the high predictive ability, offers a useful alternative to the costly and time-consuming experiments determining the bioactivity of peptide drugs.  相似文献   

13.
The molecular structure and conformation of nitrobenzene has been reinvestigated by gas-phase electron diffraction (GED), combined analysis of GED and microwave (MW) spectroscopic data, and quantum chemical calculations. The equilibrium r e structure of nitrobenzene was determined by a joint analysis of the GED data and rotational constants taken from the literature. The necessary anharmonic vibrational corrections to the internuclear distances (r e ? r a) and to rotational constants (B e (i)  ? B 0 (i) ) were calculated from the B3LYP/cc-pVTZ quadratic and cubic force fields. A combined analysis of GED and MW data led to following structural parameters (r e) of planar nitrobenzene (the total estimated uncertainties are in parentheses): r(C–C)av = 1.391(3) Å, r(C–N) = 1.468(4) Å, r(N–O) = 1.223(2) Å, r(C–H)av = 1.071(3) Å, \({\angle}\)C2–C1–C6 = 123.5(6)°, \({\angle}\)C1–C2–C3 = 117.8(3)°, \({\angle}\)C2–C3–C4 = 120.3(3)°, \({\angle}\)C3–C4–C5 = 120.5(6)°, \({\angle}\)C–C–N = 118.2(3)°, \({\angle}\)C–N–O = 117.9(2)°, \({\angle}\)O–N–O = 124.2(4)°, \({\angle}\)(C–C–H)av = 120.6(20)°. These structural parameters reproduce the experimental B 0 (i) values within 0.05 MHz. The experimental results are in good agreement with the theoretical calculations. The barrier height to internal rotation of nitro group, 4.1±1.0 kcal/mol, was estimated from the GED analysis using a dynamic model. The equilibrium structure was also calculated using the experimental rotational constants for nitrobenzene isotopomers and theoretical rotation–vibration interaction constants.  相似文献   

14.
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   

15.
The paper presents the results of a theoretical study of the dynamics of nonadiabatic transitions between the ion-pair states E0 g + and D0 u + of the I2 molecule induced by collisions with the I2 molecule in the ground electronic state X0 g + . The potential energy surfaces and diabatic coupling matrix elements of electronic states were obtained using a model based on the diatomics-in-molecule approximation. Special perturbation theory for intermolecular interaction was used to show that the large transition dipole moment between the E0 g + and D0 u + states caused the appearance of additional long-range corrections, an electrostatic dipole-quadrupole correction to the diabatic coupling matrix elements and induction dipole-dipole correction to the potential energy surface. The influence of these corrections on nonadiabatic dynamics was studied at the level of the semiclassical approximation. The electrostatic correction was found to sharply increase the contribution of resonance (accompanied by minimum kinetic energy changes) vibronic transitions at large distances between the colliding molecules. The induction correction had the opposite effect because of the high transition probability at short distances. The results obtained were in qualitative agreement with experimental data. The conclusion was drawn that obtaining quantitative agreement required a more balanced inclusion of interactions at short and long distances.  相似文献   

16.
The single crystals of Rb2[(UO2)2(C2O4)2(SeO4)] · 1.33H2O were synthesized and studied by X-ray diffraction. The crystals are monoclinic, space group P21/m, Z= 2, the unit cell parameters: a = 5.6537(8), b = 18.736(3), c = 9.4535(15) Å, β = 98.440(5)°, V = 990.6(3) Å3, R 1 = 0.0506. The main structural units of the crystal are infinite layers of [(UO2)2(C2O4)2(SeO4)]2?, corresponding to the crystal chemical group A2K 2 02 B2 (A = UO 2 2+ , K02 = C2O 4 2? , B2 = SeO 4 2? ) of uranyl complexes. The uranium-containing layers are united into a three-dimensional framework through the electrostatic interactions with the outer-sphere rubidium ions and the hydrogen bonding system involving the outer-sphere water molecules.  相似文献   

17.
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   

18.
Heterometallic pivalate Co2Sm(Piv)7(2,4-Lut)2 (1) was prepared for the first time and structurally characterized at 293 and 160 K. Antiferromagnetic exchange interactions are dominant in complex 1. This compound experiences a first-order phase transition within 210–260 K. A set of thermodynamic functions was obtained for this complex (C p , H T 0 - H 180 0 , and S T 0 ), and parameters were determined for solid-phase thermolysis where samarium cobaltate SmCoO3 is the only product.  相似文献   

19.
The work proposes a method to predict changes in the heat capacity of the liquid—vapor phase transition Δ l g Ср 0 (298.2) based on modified Randi? indices for alkanes and acyclic oxygen-containing compounds: alcohols, aldehydes, ketones, ethers, and esters. Based on the obtained Δ l g Ср 0 (298.2) values heat capacities in the liquid phase Cp liq 0 (298.2) are calculated for the compounds under study.  相似文献   

20.
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   

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