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 328 毫秒
1.
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   

2.
Molecular properties are computed as responses to perturbations (energy derivatives) in coupled-cluster (CC)/many-body perturbation theory (MBPT) models. Here, the CC/MBPT energy derivative with respect to a general two-electron (2-e) perturbation is assembled from gradient theory for 2-e property evaluation, including the electron repulsion energy. The correlation energy (?E) is shown to be the sum of response kinetic (?T), electron–nuclear attraction (?V), and electron repulsion (?V ee ) energies. Thus, evaluation of total V ee for energy component analysis is simple: For total energy (E), total 1-e responses T and V, and nuclear–nuclear repulsion energy (V NN ), V ee  = E ? V NN  ? T ? V is the true 2-e response value. Component energy analysis is illustrated in an assessment of steric repulsion in ethane’s rotational barrier. Earlier SCF-based results (Bader et al. in J Am Chem Soc 112:6530, 1990) are corroborated: The higher-energy eclipsed geometry is favored versus staggered in the two repulsion energies (V NN and V ee ), while decisively disfavored in electron–nuclear attraction energy (V). Our best quality calculations (CCSD/cc-pVQZ) attain practical Virial Theorem compliance (i.e., agreement among the kinetic energy, potential energy, and total energy representations) in assigning 2.70 ± 0.06 to the barrier height; ?195.80 kcal/mol is assigned to the drop in “steric” repulsion upon going to the eclipsed geometry. Steric repulsion is not responsible for any fraction of the ~3 kcal/mol barrier.  相似文献   

3.
The temperature dependence of heat capacity C° p = f(T) of crystalline arsenate Mg0.5Zr2(AsO4)3 was studied by precision adiabatic vacuum and differential scanning calorimetry in the temperature range 8?670 K. The standard thermodynamic functions C° p (T), H°(T)–H°(0), S°(T), and G°(T)–H°(0) of the arsenate for the range from Т → 0 to 670 K and the standard formation entropy at Т = 298.15 K were calculated from the obtained experimental data. Based on the low-temperature capacity data (30–50 K) the fractal dimension D of the arsenate was determined, and the topology of its structure was characterized. The results were compared with the thermodynamic data for the structurally related crystalline phosphates M0.5Zr2(PO4)3 (M = Mg, Ca, Sr, Ba, Ni) and arsenate NaZr2(AsO4)3.  相似文献   

4.
The first vertical electron affinities EA of 13 series of molecules and free radicals D(X i ) n are related to the inductive (σ I ), resonance (σ R ? ), and polarization (σα) parameters of substituents X i by the dependences EA = EA H + aΣσ I + bΣσ R/? + cΣσα: In radical anions D(X i ) n , compared to radical cations D(X i ) n , the polarization interaction is weaker or similar in magnitude but has an opposite sign. The previously unknown resonance parameters σ R ? of substituents SiMe3 and CH2SiMe3 bound to the radical anion center H2C=CH were calculated.  相似文献   

5.
Relations for the apparent molar heat capacity ?c of urea in an aqueous solution depending on the molality m and temperature were obtained. A transition to the relations ?c(m,T) for D2O-(ND2)2CO and T2O-(NT2)2CO systems was effected by temperature scaling. At low temperatures, the isotherms of the molar heat capacity C p(m) of the protium and deuterium systems have minima shifted to more dilute solutions at elevated temperatures. At m = 1, C p of a solution does not depend on temperature in both systems. The dependences C p(T) also have minima at constant concentrations. The temperature of the minimum heat capacity is most effectively lowered by small additions of urea. For m = 0.25, T min is 7.5 K lower than T min of pure water, and its heat capacity is 0.08 J/(mol K) higher. A transition from m = 1.5 to m = 2 lowers the temperature of the minimum heat capacity by 3.6 K; thus, the heat capacity of solutions differs by 0.02 J/(mol K) only.  相似文献   

6.
The solidification behavior of AZ Magnesium alloys in various cooling conditions was investigated using a computer-aided cooling curve thermal analysis method. In each case, the cooling curve and its first and second derivative curves have been plotted using accurate thermal analysis equipment and solidification characteristics were recognized from these curves. The cooling rates used in the present study range from 0.22 to 8.13 °C s?1. The results of thermal analysis show that the solidification parameters of AZ alloys such as nucleation temperature (T N,α), nucleation undercooling (?T N,α), recalescence undercooling (?T R,α), range of solidification temperature (?T S) and total solidification time (t f) are influenced by variation of cooling rate. Also, the effect of Al content on these parameters was studied. Microstructural evaluation was carried out to determine the correlation between the cooling rate and secondary dendrite arm spacing.  相似文献   

7.
Fine structure levels in an external magnetic field and angular dependences of resonance magnetic fields on the direction of an external magnetic field were calculated for two axially symmetrical quintet dinitrenes with the zero-field splitting parameters D q = 0.260 cm?1, E q = 0.000 and D q = 0.243 cm?1, E q = 0.003 cm?1. The EPR spectra of such dinitrenes contained lines of only three xy transitions (xy 1, xy 2, and xy 4), two Δm s = ±2 transition lines between the W ?2 and W 0 sublevels, and three additional lines from noncanonically oriented molecules whose magnetic axis Z made an angle of 12°–16° or 52°–54° with an external magnetic field.  相似文献   

8.
The phase diagram of the ternary reciprocal system K, Pb∥Cl, WO4 was studied for the first time by the calculation–experimental method and differential thermal and X-ray powder diffraction analyses. Chemical interactions between components were described, metathesis and complexation reactions were revealed, and the coordinates of binary and ternary eutectics were found (mol %): e4(410°C, 48% KCl, 52% PbCl2), e5(424°C, 23% KCl, 77% PbCl2), P(490°C, 63.5% KCl, 36.5% PbCl2), e6(487°C, 91% PbCl2, 9% PbWO4), e7(428°C, 30.5% KCl, 60.5% PbCl2, 9% PbWO4) (eutectic in the stable section D2–PbWO4), e8(650°C, 80% KCl, 20% PbWO4), e9(650°C, 70% KCl, 15% K2WO4, 15% PbWO4) (binary eutectic in the stable section D1–KCl), E1(620°C, 59% KCl, 34% K2WO4, 7% PbWO4), E2(640°C, 75% KCl, 5% K2WO4, 20% PbWO4), E3(400°C, 46% KCl, 6% PbWO4, 48% PbCl2), E4(410°C, 21% KCl, 9% PbWO4, 70% PbCl2), and Pо(468°C, 56% KCl, 10% PbWO4, 34% PbCl2).  相似文献   

9.
Sodium carboxymethyl cellulose (SCMC) with different degrees of substitution (DS) possesses structural characteristics and physicochemical properties that are important in broad areas of industrial applications. This reported work investigated the structural characteristics, including the effective length (L ef), the radius of gyration (R g), and the hydrodynamic radius (R H), and the physicochemical properties, including intrinsic viscosity ([η]) and salt tolerance, of SCMC with a DS more than 1.0 in NaCl solution using molecular dynamics (MD) simulations. In the MD simulations, the DS of SCMC varied from 1.2 to 2.8, and the NaCl concentration varied from 0 to 1.4 mol/L. MD simulation results showed that with the increment of NaCl concentration, the L ef (or R g or R H) of SCMC decreased; with the increment of the DS, the L ef of SCMC increased. Also, the variation tendency of [η] in the NaCl solution was consistent with its L ef (or R g or R H). It was noted that the salt tolerance (represented by D) of SCMC increased as the DS increased. In addition, the sharp variation of the D value of SCMC occurred in the range of DS of 1.6 to 2.0, which agreed with the reported experimental results. Radial distribution function analyses showed that the Na+ cations had a stronger interaction with the carboxyl groups in SCMC with lower DS when it was present in a salt solution of higher concentration, which also reasonably explained the variation of L ef, R g, R H, [η], and D of SCMC in NaCl solution.  相似文献   

10.
The applications of the Sand equation in potentiometry of electrode and membrane systems for precise measurements of the transition time (τ) have been determined. An approach was suggested for choosing the diffusion coefficient of electrolyte (D) in the case when the concentration changes from its value in the agitated solution (where D = Db) to the nearly zero value at the surface (D = D0 corresponds to an infinitely dilute solution), Db and D0 being substantially different. The Nernst–Planck–Poisson nonstationary equations were numerically solved in a one-dimensional system including an ion-exchange membrane and two adjacent diffusion layers (for the electrode–solution system, the result is a particular case). An effective value Def was found, whose substitution in the Sand equation gave τ identical to that obtained by numerical solution. The neglect of the concentration dependence D(с) can lead to a nonadequate determination of the ion transport numbers in the membrane.  相似文献   

11.
The O?H bond dissociation energy (D O?H) has been determined for eight alkylseleno-substituted phenols, one alkyltelluro-substituted phenol, and one alkyltelluro-substituted pyridinol. D O?H has been estimated by the intersecting-parabolas method from kinetic data using five reference compounds: α-tocopherol (D O?H = 330.0 kJ/mol), 3,5-di-tert-butyl-4-methoxyphenol (D O?H = 347.6 kJ/mol), 4-methylphenol (D O?H = 361.6 kJ/mol), 2,6-di-tert-butyl-4-methylthiophenol (D O?H = 336.3 kJ/mol), and 2,6-di-ter-tbutyl-4-methylphenol (D O?H = 338.0 kJ/mol). The following D O?H values (kJ/mol) have been obtained: 335.9 for 2,5,7,8-tetramethyl-2-phytyl-6-hydroxy-3,4-dihydro-2H-1-benzoselenopyran, 342.6 for 2-methyl-5-hydroxy-2,3-dihydrobenzoselenophene, 333.5 for 2,4,6,7-tetramethyl-5-hydroxy-2,3-dihydrobenzoselenophene, 339.4 for 2-tert-butyl-4-methoxy-6-octylselenophenol, 357.9 for dodecyl 3-(4-hydroxyphenyl) propyl selenide, 348.5 for dodecyl 3-(3,5-dimethyl-4-hydroxyphenyl)propyl selenide, 350.9 for dodecyl 3-(3-tert-butyl-4-hydroxyphenyl)propyl selenide, 338.0 for dodecyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl) propyl selenide, 343.0 for 2,6-di-tert-butyl-4-(tellurobutyl-4′-phenoxy)phenol, and 338.8 for 6-octyltelluro-3-pyridinol. The stabilization energies of phenoxyl radicals containing R substituents (X = O, S, Se, Te) have been compared.  相似文献   

12.
The temperature dependence of the heat capacity of triphenylantimony dibenzoate Ph3Sb(OC(O)Ph)2 is studied in the range of 6–480 K by means of precision adiabatic vacuum calorimetry and differential scanning calorimetry. The melting of the compound is observed in this temperature range, and its standard thermodynamic characteristics are identified and analyzed. Ph3Sb(OC(O)Ph)2 is obtained in a metastable amorphous state in a calorimeter. The standard thermodynamic functions of Ph3Sb(OC(O)Ph)2 in the crystalline and liquid states are calculated from the obtained experimental data: Cp°(T), H°(T)–H°(0), S°(T), and G°(T)–H°(0) for the region from T → 0 to 480 K. The standard entropy of formation of the compound in the crystalline state at T = 298.15 K is determined. Multifractal processing of the low-temperature (T < 50 K) heat capacity of the compound is performed. It is concluded that the structure of the compound has a planar chain topology.  相似文献   

13.
The formulas for calculation of the number of atoms in nanoparticles with symmetry group D 6h are reported. The numbers of atoms are determined by six structurally invariant numbers and the “quantum number” of the group order n. Eight classes of nanostructures with symmetry group D 6h are revealed: C ? + 12z , where z = 0, 1, 2, …, and C ? is C 2, C 6, C 8, or C 14. The sum rule for the coordination numbers of all atoms of subshells related to symmetry elements is established. Two-dimensional nanoparticles are considered.  相似文献   

14.
The formation of chloro- and azidocomplexes of VO2+(IV) is investigated in acetonitrile (AN), propanediol-1,2-carbonate (PDC), trimethyl phosphate (TMP) by spectrophotometric, potentiometric and conductometric methods. The following coordination forms are indicated: [VOCl]+ inAN, PDC andDMSO), [VOCl2] (inAN, PDC andTMP), [VOCl3]? (inPDC andTMP[?]), [VOCl4]2? (inAN, PDC andTMP); [VON3]+ (inAN, PDC andDMSO), [VO(N3)2] (inAN, PDC, TMP andDMSO), [VO(N3)2+n]n? (inAN, PDC, TMP andDMSO). The results are interpreted by the donor numbers and sterical properties of the solvent molecules.  相似文献   

15.
The chloro systems of Mn2+ and Fe3+ were investigated in acetonitrile (AN), propanediole-1,2-carbonate (PDC) and trimethylphosphate (TMP) by spectrophotometric, potentiometric and conductometric methods. The following coordination forms seem to be present: MnCl2 (inAN, PDC andTMP), [MnCl4]2? (tetrahedral inAN, octahedral inTMP), [MnCl6]4? (octahedral inPDC andTMP); [FeCl]2+ (tetrahedral inAN andPDC), [FeCl2]+ (tetrahedral inTMP), FeCl3 (tetrahedral inAN andPDC), [FeCl4]? (tetrahedral inAN, PDC andTMP).  相似文献   

16.
The present study is a comparative study of three equations, namely the Clausius–Clapeyron, Van’t Hoff and Hildebrand (to calculate crystal–liquid fugacity ratio (CLFR) of drug compounds), to select the best model in predicting the intestinal absorption and develop a new classification system based on dose number (D o) and CLFR. The required thermodynamic parameters [melting point, enthalpy of fusion (ΔH m) and the differential molar heat capacity (?C pm)] were experimentally obtained by differential scanning calorimetry. Pharmacokinetic data [the human intestinal absorption (F a) and apparent permeability of Caco-2 (P app _Caco-2)] and D o were obtained from the literature. The highest value of CLFR was found for diclofenac with the value of 88.78, 87.29, and 87.84 mol% from Clausius–Clapeyron, Van’t Hoff, and Hildebrand approaches, respectively. The lowest CLFR value was seen for memantine with the value of 14.3 × 10?17 and 26 × 10?12 mol% from Van’t Hoff and Hildebrand equations, respectively. Statistical comparison with the Wilcoxon signed rank test showed that the CLFR values calculated by three equations are different. CLFR values of more than 1 mol% correspond to the complete intestinal absorption (F a). There was a sigmoidal dependency between CLFR and F a, similar to the dependency between P app _Caco-2 and F a. In these modeling, the excellent correlations were obtained in all three models as evidenced by a good coefficient of determination (r 2 ) without a significant difference in the average absolute error. A new classification system from Hildebrand model based on D o and CLFR was developed and was in agreement with the biopharmaceutics classification system (70.5%) and the biopharmaceutical drug disposition system (65.6%). This modeling approach can be a valuable tool for scientists as an alternative for intestinal permeability in the biopharmaceutical classification system to develop new oral drugs. The CLFR obtained from Hildebrand model is also more convenient than the Clausius Clapeyron model, because the former does not need to calculate ?C pm (difficult step in calculating CLFR) for drug compounds. This new classification can help to develop the new drug product in industrial and academic research, without necessary in vivo experiments.  相似文献   

17.
Addition of chloride ions to the hexasolvated ions of Ti(III), V(III) and Cr(III) in propandiol-1,2-carbonate (PDC) and trimethylphosphate (TMP) may lead to the following complexes: [TiCl]2+ (inPDC), [TiCl2]+ (inTMP), TiCl3 (inPDC andTMP, low solubility inTMP), [TiCl4]? (inPDC andTMP?), [TiCl6]3? (inPDC); [VCl]2+ (inPDC andTMP), VCl3 (inPDC andTMP, low solubility inTMP), [VCl4]? (inPDC); [CrCl]2+ (inTMP), [CrCl2]+ (inPDC), CrCl3 (inPDC andTMP), [CrCl4]? (inPDC).  相似文献   

18.
A new nonparametric scaling equation of state is suggested. The equation correctly describes the p-ρ-T data and heat capacities of liquids close to the critical vaporization points. It was obtained with the use of the S spinodal and the mixing of scaling fields as a first approximation (asymmetric and nonasymptotic terms were ignored). The new equation was used to approximate the data on 4He, C2H4, and H2O in the critical region. The results showed that it correctly described the critical behavior of thermodynamic functions, including isochoric heat capacity, not only in the asymptotic but also over a fairly wide density region at the critical point. The suggested equation of state describes the p-ρ-T data with the same error as the Schofield parametric equation of state. The new equation, however, better reproduces the behavior of heat capacities and is much simpler to use. As distinct from the Schofield equation, the new equation, like classic equations of state, allows the spinodal to be determined from the (?p/?v) T = 0 condition at T < T c .  相似文献   

19.
In this work, iron nanoparticles were impregnated onto a commercial activated carbon surface to produce a novel adsorbent called iron-activated carbon nanocomposite (I-AC). Commercial activated carbon (CAC) and I-AC were used for vanadium separation in a fixed-bed column. The effects of various operating parameters such as inlet vanadium ion concentration, adsorbent dose and volumetric flow rate on vanadium separation performance of CAC were investigated. The performance of both adsorbents was compared in three adsorption/desorption cycles. The experimental breakthrough curves of vanadium ions in the fixed-bed column were modeled using the film-pore-surface diffusion model (FPSDM). The four mass transfer parameters characterizing this model, namely the external mass-transfer coefficient (k f ), pore and surface diffusion coefficients (D p and D s ), and axial dispersion coefficient (D L ) were evaluated through the model. Modelling and experimental results showed that the I-AC nanocomposite has a better performance for vanadium separation in comparison to AC. Sensitivity analysis on the FPSDM showed that the pore and surface diffusion, external mass transfer and axial dispersion play a significant role in vanadium separation using the I-AC. On the other hand, surface diffusion resulted to be relatively less important when CAC was used.  相似文献   

20.
The heat capacity of a glassy third-generation poly(phenylene-pyridyl) dendron decorated with dodecyl groups is studied for the first time via high-precision adiabatic vacuum and differential scanning calorimetry in the temperature range of 6 to 520 K. The standard thermodynamic functions (molar heat capacity Cp°, enthalpy H°(T), entropy S°(T), and Gibbs energy G°(T)-H°(0)) in the range of T → 0 to 480 K, and the entropy of formation at 298.15 K, are calculated on the basis of the obtained data. The thermodynamic properties of the dendron and the corresponding third-generation poly(phenylene-pyridyl) dendrimer studied earlier are compared.  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