首页 | 官方网站   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.
Dilatometric measurements of excess volume VE and ultrasonic speed u have been carried out for mixtures of mono-, di-, tri- and tetra(ethylene glycol)s in pyrrolidin-2-one (PY) over the whole mole fraction range at 303.15 K. In the mixture of PY and monoethylene glycol, the VE is positive except for slight negative variation at the high mole fraction of PY. The other three mixtures PY + di-, + tri- and + tetra(ethylene glycol)s show negative VE over the entire composition range in the order di-u with increase in the mole fraction of PY in the case of monoethylene glycol while for other three systems u rises. From these measurements, partial molar quantities ViE and KS,iE have been calculated and analysed. Estimates of isentropic molar quantity KS equal to −(∂V/∂p)S and its excess counterpart KSE have also been computed. The KSE is positive for mono-, and negative for all the other mixtures over the whole composition range.  相似文献   

2.
Densities of {(1−x)CH3(CH2)n−1OH + xCH3CN} for n=1, 2, 3 or 4 have been determined as a function of composition at 288.15, 293.15, 298.15 and 303.15 K at atmospheric pressure using a vibrating-tube densimeter (Anton Paar DMA 4500, resolution 1×10−5 g cm−3). Excess molar volumes were calculated. The VmE values were negative for acetonitrile–methanol mixtures and sigmoid for acetonitrile–alkanols (C2–C4) mixtures over the complete mole fraction range. VmE values increase in a positive direction with increase in chain length of the alkanols and with the temperature. The Extended Real Associated Solution Model (ERAS-Model) calculations allowing for self-association for the alkanols and complex formation between acetonitrile and alkanols have been used to correlate experimental data. The model is able to reproduce the asymmetrical VmE behavior of the studied systems, although agreement between theoretical and experimental values is less satisfactory for some concentration ranges.  相似文献   

3.
The dielectric permittivity tensor components, εII and ε, in the nematic phase of 6CB (4-n-hexyl-4'-cyanobiphenyl) were measured in the pressure range 0.1-130 MPa and the temperature range 12-58°C. The dielectric anisotropy, Δε(p, V, T) = εII - ε, was analysed in isothermal, isobaric and isochoric conditions taking into account the pVT data and the well known Maier and Meier equation. On that basis the nematic order parameter S(p, V, T) was determined. This was used to calculate the parameter Γ relating the interaction potential with the volume (density). Its value Γ = 4.1 agrees very well with other estimates.  相似文献   

4.
Some aspects of the theory of LASIN (laser assisted surface ion neutralization) are discussed, with emphasis on the physical origins of the so-called double-peak structures found in some calculations of the charge-transfer (neutralization) probability, P, as a function of the laser frequency η. These two peaks have been called the first peak at η ≈ ηm = Om (in a.u.), where o(m) is the electronic energy level of the ion/atom (middle of the solid's valence band) and the second peak, a much larger peak at η ≈ 1.3 ηm, respectively.

We show that these double-peak structures are all special cases of multiple-peak structures which result from quantum interference effects, and that, in fact, the second peak is to be regarded as the main resonance peak. This result is interesting in itself, because it is the first peak which has heretofore been considered the main resonance peak.

To simplify the discussion, a two-level model is adapted, which represents the solid valence band by a single level at m. Clarification of the physical reason for the multiple peaks is based on the semiclassical theory of nonadiabatic transitions, in which the peaks are due to the phase difference between the two adiabatic paths that arise from the diagonalization of the two-level hamiltonian.

With the electronic hopping potential modelled by V(t) = Vosech(λt), and the laser potential by W(t) = Wosech(λt) cos(πt + δ), in the usual notation, an approximate analytical expression for P(η) is presented for the case Wo/Vo < 1, which covers most of the previous treatments, and is in good agreement with the exact results.  相似文献   


5.
The excess molar volumes VmE {x(CH3OH or CH3CH2OH or CH3(CH2)2OH or CH3CH(OH)CH3 + (1 - x){CH3(CH2)2}2O or CH3C(CH3)2OCH3 or CH3CH2C(CH3)2OCH3} have been calculated from measured values of density over the whole composition range at the temperature 298.15 K in order to investigate OH … O specific interactions. The results are explained in terms of the strong self-association of the alkanols, the specific interaction between the alkanol, and the ether molecules and packing effects upon mixing. The experimental Vmh results presented here, together with the previously reported data for the molar excess enthalpy HmE, has been used to test the Extended Real Associated Solution (ERAS) model.  相似文献   

6.
This paper reports excess molar enthalpies, HmE, and excess molar volumes, VmE, of the binary systems {propyl propanoate + o-xylene}, {propyl propanoate + m-xylene} and {propyl propanoate + p-xylene} at the temperature 298.15 K and atmospheric pressure, over the whole composition range. VmE was calculated from the experimental measurement of the corresponding densities, while HmE was measured directly. The excess magnitudes were correlated to a Redlich-Kister type equation. Finally, we will discuss the results of the three mixtures studied here and by comparison with other binary systems containing propyl propanoate and a benzene-based compound previously published.  相似文献   

7.
Molar excess enthalpies HmE, isobaric heat capacities CP,mE, volumes VmE and isothermal compressibilities κTE for the 1,3-dioxane(3DX) + cyclohexane mixture were measured at 298.15 K, in order to compare to those of the 1,4-dioxane(4DX) + cyclohexane mixture. HmE is endothermic and the maximum value about 1.5 kJ mol−1 at x ≈ 0.45, and lower than that of the 4DX mixture by about 80 J mol−1. VmE is positive over the whole concentration and the maximum value is about 0.85 cm3 mol−1 at x ≈ 0.45, and lower than that of the 4DX mixture. The above results suggest the energetic unstabilization, resulting in the volume expansion in the mixture. CP,mE shows the characteristic W-shaped concentration dependence, which has maximum at x ≈ 0.45 and two minima at x ≈ 0.1 and 0.9. The maximum CP,mE value for 3DX mixture shifts toward the positive side, compared to that of 4DX mixture. κTE were estimated from speeds of sound, densities, thermal expansion coefficients and isobaric heat capacities of the pure component liquids and the mixtures. The κTE result shows the positive concentration dependence over the whole composition range. The 3DX mixture has the similar thermodynamic properties to the 4DX mixture, despite that 4DX is the nonpolar solvent and 3DX is the dipolar liquid. this means that there exists the local dipolar interaction between 4DX molecules, and the prevalence of “microheterogeneity” in the both mixtures.  相似文献   

8.
The Kováts coefficients, Kc,Z, of a stationary phase and the solute's molecular structural coefficients, Sc,i, depend both on the specific retention volume Vg, of a solute or homologous series and on the “log-plot” slope, b, of a chromatographic column. In view of this dependence, the feasibility of predicting Vg in three instances was investigated: (a) Vg prediction of any n-alkane from Kc,Z and retention data of n-decane; (b) Vg prediction of any solute from the temperature dependence of the above parameters and (c) Vg prediction of any term of a homologous series from the correlations of the Sc increments, ΔSc, with the organic structural function. The possibilities of the method are evaluated in the light of the analysis of the deviations of the predicted Vg values from the measured values.  相似文献   

9.
Even for such simple mixtures as (argon+methane), the excess enthalpy HEm and the excess volume VEm in the near critical region are about two orders of magnitude higher than for the liquid mixture at low temperatures and pressures near ambient conditions. Mixtures for which the critical temperatures are close together, and for which the critical pressures are far apart, have similar HEm (x,p,T) and VEm (x,p,T) surfaces, and near critical isotherms show double maxima in the supercritical fluid region. Mixtures for which the critical pressures are close together, and the critical temperatures are far apart, also have similar HEm (x,p,T) and VEm (x,p,T) surfaces, but isobars on the surfaces are ‘S’ shaped. The shapes of these near-critical excess-function surfaces can be understood from an inspection of the enthalpy, or residual enthalpy curves of the mixture and of the pure components. Examples of both are given. Attention is drawn to the large value that these excess functions can have close to a pure component critical point.  相似文献   

10.
The structure of the complex [Ni(hmt)(NCS)2(H2O)2]n, assembled by hexamethylenetetramine (hmt) and octahedral Ni(II), is reported. Crystal data: Fw 351.07, a=9.885(10) Å, b=12.06(1) Å, c=12.505(8) Å, β=114.41(4)°, V=1357(1) Å3, Z=4, space group=C2/c, T=173 K, λ(MoK)=0.71070 Å, ρcalc=1.718 gcm−1, μ=17.44 cm−1, R=0.099, Rw=0.145. The tetrahedral assembling template effect of the hmt molecule is completed by two coordination bonds and two hydrogen interactions. The UV–vis absorption spectrum of this complex [Ni(hmt)(NCS)2(H2O)2]n with a two-dimensional network is determined in the range of 5000–35000 cm−1 at room temperature. The observed spectrum is discussed and explained perfectly by the scaling radial theory proposed by us. The two-dimensional structure has no apparent effects on the d–d transitions of the central Ni(II) ion. The IR spectrum and the GT curve of the complex were also measured and clearly reflect its structural properties.  相似文献   

11.
A calorimetric study was performed for adducts of general formula CdBr2·nL (n=1 and 2; L=ethyleneurea (eu) and propyleneurea (pu)). The standard molar reaction enthalpy in condensed phase: CdBr2(c)+nL(c)=CdBr2·nL(c); ΔrHmθ, were obtained by reaction–solution calorimetry, to give the following values for mono- and bis-adducts: −19.54 and −34.59; −7.77 and −19.05 kJ mol−1 for eu and pu adducts, respectively. Decomposition (ΔDHmθ) and lattice (ΔMHmθ) enthalpies, as well as the mean cadmium---oxygen bond dissociation enthalpy, DCd---O, were calculated for all adducts.  相似文献   

12.
CaRgn+ (Rg=He, Ne, Ar) complexes with n=1–4, are investigated by performing using the B3LYP/6-311+G (3df) density functional theory calculations. The CaHen+ (n=1–4) complexes are found to be stable. In the case of CaNen+ and CaArn+, stable structures and stationary point were found only for n=1 and 2. For n=3 in the C3V and the D3h point group as well as for n=4 in the Td (tetrahedral) point group a saddle point (imaginary frequency) is observed and global minimum could be obtained along the potential energy surface.  相似文献   

13.
The neutral nitrogen-bidentate ligand, diphenylbis(3,5-dimethylpyrazol-1-yl)methane, Ph2CPz′2, can readily be obtained by the reaction of Ph2CCl2 with excess HPz′ in a mixed-solvent system of toluene and triethylamine. It reacts with [Mo(CO)6] in 1,2-dimethoxyethane to give the η2-arene complex, [Mo(Ph2CPz′2)(CO)3] (1). This η2-ligation appears to stabilize the coordination of Ph2CPz′ 2 in forming [Mo(Ph2CPz′2)(CO)2(N2C6H4NO2-p)][BPh4] (2) and [Mo(Ph2CPz′2)(CO)2(N2Ph)] [BF4] (3) from the reaction of 1 with the appropriate diazonium salt but the stabilization seems not strong enough when [Mo{P(OMe)3} 3(CO)3] is formed from the reaction of 1 with P(OMe)3. The solid-state structures of 1 and 3 have been determined by X-ray crystallography: 1-CH2Cl2, monoclinic, P21/n, a = 11.814(3), b = 11.7929(12), c = 19.46 0(6) Å, β = 95.605(24)°, V = 2698.2(11) Å3, Z = 4, Dcalc = 1.530 g/cm3 , R = 0.044, Rw = 0.036 based on 3218 reflections with I > 2σ(I); 2 (3)-1/2 hexane-1/2 CH3OH-1/2 H2O-1 CH2Cl2, monoclinic, C2/c, a = 41.766(10), b = 20.518(4), c = 16.784(3) Å, β = 101.871(18)°, V = 14076(5) Å3, Z = 8, Dcalc = 1.457 g/cm3, R = 0.064, Rw = 0.059 based on 5865 reflections with I > 2σ(I). Two independent cations were found in the asymmetric unit of the crystals of 3. The average distance between the Mo and the two η2-ligated carbon atoms is 2.574 Å in 1 and 2.581 and 2.608 Å in 3. The unfavourable disposition of the η2-phenyl group with respect to the metal centre in 3 and the rigidity of the η2-arene ligation excludes the possibility of any appreciable agostic C---H → Mo interaction.  相似文献   

14.
The synthesis and reactivity of {(η5-C5H4SiMe3)2Ti(CCSiMe3)2} MCl2 (M = Fe: 3a; M = Co: 3b; M = Ni: 3c) is described. The complexes 3 are accessible by the reaction of (η5-C5H4SiMe3) 2Ti(CSiMe3)2 (1) with equimolar amounts of MCl2 (2) (M = Fe, Co, Ni). 3a reacts with the organic chelat ligands 2,2′-dipyridyl (dipy) (4a) or 1,10-phenanthroline (phen) (4b) in THF at 25°C to afford in quantitative yields (η5-C5H4SiMe3)2Ti(CSiMe3)2 (1) and [Fe(dipy)2]Cl2 (5a) or [Fe(phen)2]Cl2 (5b). 1/n[CuIHal]n (6) or 1/n[AgIHal]n (7) (Hal = Cl, Br) react with {(η5 -C5H4SiMe3)2Ti(CCSiMe3)2}FeCl2 (3a), by replacement of the FeCl2 building block in 3a, to yield the compounds {(η5-C5H4SiMe3)2Ti(C CSiMe3)2}CuIHal (8) or {(η5-C5H4SiMe3)2Ti(CSiMe3)2}AgIHal (9) (Hal = Cl, Br), respectively. In 8 and 9 each of the two Me3SiCC-units is η2-coordinated to monomeric CuI Hal or AgIHal moieties. Compounds 8 and 9 can also be synthesized by the reaction of (η5-C5H4SiMe3)2 Ti(CSiMe3)2 (1) with 1/n[CuIHal]n (6) or 1/n [AgIHal]n (7) in excellent yields. All new compounds have been characterized by analytical and spectroscopic data (IR, 1H-NMR, MS). The magnetic moments of compounds 3 were measured.  相似文献   

15.
Excess molar enthalpies HE and excess molar volumes VE have been measured, as a function of mole fraction x1, at 298.15 K and atmospheric pressure for the five liquid mixtures (x11,4-C6H4F2 + x2n-ClH2l+2), l = 7, 8, 10, 12 and 16. In addition, HE and excess molar heat capacities CPE at constant pressure have been determined for the two liquid mixtures (x1C6F6 + x2n-ClH2l+2), l = 7 and 14, at the same temperature and pressure. The instruments used were flow microcalorimeters of the Picker design (the HE version was equipped with separators) and a vibrating-tube densimeter, respectively.

The excess enthalpies of the five difluorobenzene mixtures are all positive and quite large; they increase with increasing chain length l of the n-alkane from HE(x1 = 0.5)/(J mol−1) = 1050 for l = 7 to 1359 for l = 16. The corresponding excess volumes VE are all positive and also increase with increasing l: VE(x1 = 0.5)/(cm3 mol−1) = 0.650 for l = 7 and 1.080 for l = 16. Interestingly, the excess enthalphies of the corresponding mixtures with hexafluorobenzene are only about 5% larger, whereas the excess volumes of (x1C6F6 + x2n-ClH2l+2) are roughly twice as large as those of their counterparts in the series containing 1,4-C6H4F2. Specifically, at 298.15 K HE(x1 = 0.5)/(J mol−1) = 1119 for (x1C6F6 + x2n-C7H16) and 1324 for (x1C6F6 + x2n-C14H30), and for the same mixtures VE(x1 = 0.5)/(cm3 mol−1) = 1.882 and 2.093, respectively. The excess heat capacities for both systems are negative and of about the same magnitude as the excess heat capacities of mixtures of fluorobenzene with the same n-alkanes (Roux et al., 1984): CPE(x1 = 0.5)/(J K−1 mol−1) = −1.18 for (x1C6F6 + x2n-C7H16), and −2.25 for (x1C6F6 + x2n-C14H30). The curve CPE vs. (x1 for x1C6F6 + x2n-C14H30) shows a sort of “hump” for x1 0.5, which is presumed to indicate emerging W-shape composition dependence at lower temperatures.  相似文献   


16.
The siloxyanilines o-Me3SiOC6H4NH2 (1) and p-RMe2SiOC6H4NH2 (R=H (2); R=Me (3)), and their N-silylated derivatives p-Me3SiOC6H4NHSiMe3 (4) and p-Me3SiOC6H4N(SiMe3)2 (5) have been prepared from ortho- or para-aminophenol and used in the synthesis of imido complexes. Thus, binuclear [{Ti(η5-C5H5)Cl}{μ-NC6H4(p-OSiMe3)}]2 (6) and mononuclear [TiCl2{NC6H4(p-OSiMe3)}(py)3] (7) imido complexes have been obtained from the reaction of 3 and [Ti(η5-C5H5)Cl3] or [TiCl2(NtBu)(py)3], respectively. In contrast, the reaction of 1 with TiCl4 and tBupy affords the titanocycle [TiCl2{OC6H4(o-NH)---N,O}(tBupy)2] (8). Compound 5 has also been used to prepare the niobium imide complex [NbCl3{NC6H4(p-OSiMe3)}(MeCN)2] (9), by its reaction with NbCl5 in CH3CN. These findings have been applied to the synthesis of polynuclear systems. Thus, chlorocarbosilane Si[CH2CH2CH2Si(Me)2Cl]4 (CS–Cl) has been functionalized with the ortho- and para-aminophenoxy groups to give 10 and 11, respectively. The use of 11 has allowed the formation of the tetranuclear compound 12. Attempts to synthesize terminal imido titanium complexes from 10 and TiCl4 in the presence of tBupy and Et3N, give complex 8 and carbosilane CS–Cl.  相似文献   

17.
De-Dong Wu  Thomas C. W. Mak 《Polyhedron》1994,13(24):3333-3339
Two polymeric mercury(II) halide adducts of an olefinic double betaine, cis-(p-Me2NC5H4N+)2C2(COO)2 (L), have been prepared and characterized by X-ray crystallography. [{Hg2L2Cl4·6HgCl2}n] (1) crystallizes in the monoclinic space group C2/c with Z = 4, and [{Hg2L2Br4·HgBr2}n] (2) in the triclinic space group P with Z = 1. Complexes 1 and 2 are structurally similar, being composed of centrosymmetric fourteen-membered rings and nearly linear HgX2 (X = Cl, Br) moieties that are further inter-linked by weak HgX [HgCl = 2.930–3.136(9) Å, HgBr = 3.057–3.310(6) Å] and HgO [2.64, 2.75(3) Å] bonds to generate a two-dimensional polymeric network.  相似文献   

18.
The methylene-bridged, mixed-chalogen compounds Fe2(CO)6(μ-SeCH2Te) (1) and Fe2(CO)6(μ-SCH2Te) (3) have been synthesised from the room temperature reaction of diazomethane with Fe2(CO)6(μ-SeTe) and Fe2(CO)6(μ-STe), respectively. Compounds 1 and 3 have been characterised by IR, 1H, 13C, 77Se and 125Te NMR spectroscopy. The structure of 1 has been elucidated by X-ray crystallography. The crystalsare monoclinic,space group P21/n, A = 6.695(2), B = 13.993(5), C = 14.007(4)Å, β = 103.03(2)°, V = 1278(7) Å3, Z = 4, Dc = 2.599 g cm−3 and R = 0.030 (Rw = 0.047).  相似文献   

19.
Gas electron diffraction is applied to determine the geometric parameters of the silacyclobutane molecule using a dynamic model where the ring puckering was treated as a large amplitude motion. The structural parameters and the parameters of the potential function were refined taking into account the relaxation of the molecular geometry estimated from ab initio calculations at the MP2/6-311+G(d, p) level of theory. The potential function has been described as V() = V0[(/e)2 − 1]2 with the following parameters V0 = 0.82 ± 0.60 kcal/mol and e = 33.5 ± 2.7°, where is a puckering angle of the ring.

The geometric parameters at the minimum V() (ra in Å, in degrees and uncertainties given as three times the standard deviations including a scale error) are: r(Si–Hax) = 1.467(96), r(Si–Heq) = 1.468(96), r(Si–C) = 1.885(2), r(C–C) = 1.571(3), r(C–H) = 1.100(3), CSiC = 77.2(9), HSiH = 108.3, SiCHeq = 123.5(16), SiCHax = 111.9(16), CC5Heq = 118.4(24), CC5Hax = 112.3(24), HC3H = 107.7, δ(HSiH) = 6.6, δ(HC3H) = 7.0, where the tilts δ, HSiH, and HC3H are estimated from ab initio constraints. The structural parameters are compared with those obtained for related compounds.  相似文献   


20.
Meloun M  Javůrek M 《Talanta》1984,31(12):1083-1094
MRFIT and MRLET, two FORTRAN computer programs, can analyse a photometric mole-ratio curve (photometric titration curve) to estimate the stability constant βpq of the predominant complex MpLq, the ligand concentration factor fL, the extrapolated absorbance Aext and the stoichiometric coefficient q (p is usually 1). MRFIT uses algorithmic minimization of a residual-square sum to reach, usually, the global minimum or the lowest of several local ones. MRLET, an ABLET system program based on LETAG, allows algorithmic and/or heuristic minimization. A local minimum described by parametric co-ordinates with a definite physical meaning might be found by the heuristic process.  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-23

京公网安备 11010802026262号