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 15 毫秒
1.
The complete second rank ordering matrix for perylene-d12 and pyrene-d10 in four different thermotropic liquid crystals has been determined over large temperature ranges, by interpreting the quadrupolar splittings observed from the deuterium N.M.R. spectra. The dipolar couplings also observed in the spectra were determined by computer simulation for both probes, allowing the commonly made assumptions of rigidity and planarity of these molecules to be tested. It is found that the perylene results are consistent with a non-rigid and non-planar structure with an average twist angle between the napthtalene units of 11.6°, whereas pyrene is rigid and planar within experimental error. It is also observed that both molecules are highly biaxial in their orientational behaviour and therefore another assumption often made especially in fluorescence measurements, that of disc-like cylindrical symmetry, is invalid. The results are interpreted in terms of a molecular field theory which predicts the variation of (Sxx - Syy) with Szz by calculating these quantities from the potential of mean torque, U(β, γ), of rigid solutes, expressed solely in terms of one adjustable parameter, the biaxiality parameter, λ. This parameter is predicted by some theories to be both temperature and solvent independent, but our results show that there is a weak temperature dependence and a stronger solvent dependence.  相似文献   

2.
The nematic substance 5CB is known from N.M.R. studies to be slightly biaxial, not in the sense that any bulk property measured in a direction at right angles to the director is liable to vary with rotation about the director, but in the sense that there are biaxial terms in the ordering matrix that describes the alignment of individual molecules; (Sxx - Syy) is non-zero as well as Szz. We show that the biaxial terms should make a significant contribution to the magnetic anisotropy Δχ(m) of 5CB, and that the magnitude and temperature dependence of this bulk property, which we have measured, can be understood if, and only if, they are taken into account. The contribution which they make to the optical birefringence term ∑ should, however, be relatively trivial. Although ∑ may in principle be affected by local field corrections of a complicated nature, which do not affect Δχ(m), a new theory presented in an Appendix to the paper suggests that these too are likely to be relatively trivial. Hence we believe that ∑ is more nearly proportional than is Δχ(m) to the principal order parameter Szz.

The paper includes unpublished data for the magnetic anisotropy and/or the principal refractive indices, ne and no, in a number of other nematics (6CB, 7CB, 8CB, 9CB, 5OCB, 6OCB, 7OCB, 8OCB, 7CCH, MBBA and PAA). Comparison between the temperature dependence of Δχ(m) and of ∑ suggests that biaxiality is present in all cyanobiphenyls, on much the same scale as in 5CB, and is not affected by the presence of an oxygen atom between the phenyl core of the molecule and its alkyl tail.  相似文献   

3.
The dependence of the Stern potential, ψ1, of glass samples on the distance between these, H, has been theoretically calculated, while taking into account the Stern isotherm and the electroneutrality equation. Comparison of the theoretical dependences ψ1(C)H→∞ with those previously experimentally obtained enables one to calculate the energy of adsorption of OH ions on glass and, further, the dependence ψ1(H). It has been shown that for pH 4–6 and CKCl = 10-2-10-5 mol/L, the value of ψ1 practically does not depend on H. The result obtained was used to calculate theoretically the ionic-electrostatic forces and to compute (from the experimental values of the interaction forces) structural forces Us(H). The dependence thus obtained, Us(H), is of exponential character.  相似文献   

4.
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   

5.
The currently used equation for solute transport in nanofiltration contains two parameters (ω and 1 – σ) that may be concentration-dependent. A force balance equation allows interpretation of these parameters (as well as Lp) in terms of distribution and friction coefficients, as was demonstrated for a neutral solute and a single 1:1 salt. It is generally assumed in model calculations that it is the distribution coefficient that determines the concentration dependence of ω and 1 – σ. This suggests that a more practically convenient form of the equation may be proposed, in which only one concentration-dependent parameter, ω, appears, while the other is replaced with the ratio of the two, A, which has the meaning of the membrane Peclét number divided by the volume flux and may be assumed to be constant. This may facilitate the analysis of flux–rejection curves and parameter evaluation including concentration dependence, which is a crucial and unavoidable step towards predictive NF modeling. The direct connection between transport parameters and distribution coefficients also suggests that experimentally measured concentration dependence may help to discriminate between different exclusion mechanisms. An approximate analysis based on the connection between A and solute–water friction shows that for presently used NF membranes and realistic fluxes the expected contribution of convection to solute flow cannot become dominant so that the limiting value for salt rejection, R = σ, cannot be reached.  相似文献   

6.
In this study we investigate two alternative pathways to compute the free energy and the entropy of small molecule association (ΔFassoc and ΔSassoc) in water. The first route (direct pathway) uses thermodynamic integration as function of the distance R between the solutes. The mean force and the mean covariance of the force with the energy in solution are calculated from molecular dynamics simulation followed by integration of these quantities with respect to the reaction coordinate R. The alternative approach examined (solvation pathway) would first remove the solutes from the solution using thermodynamic integration as function of a solvation coupling parameter λ, change the solute–solute distance in vacuo and then solvate back the solute pair at the new separation distance. The system studied was a pair of CH4 molecules in water. We investigate the influence of the CH4–water interaction strength on the obtained ΔFassoc and ΔSassoc values by changing van der Waals and Coulomb interaction and evaluated the accuracy and efficiency for the two pathways. We find that the direct route seems more suitable for the calculation of free energies of hydrophobic solutes while the solvation pathway performs better when calculating entropy changes for solutes that have a stronger interaction with the solvent.  相似文献   

7.
The behaviour of a nematic liquid crystal when it is spun about an axis orthogonal to a magnetic field is predicted to be controlled by the critical angular velocity, ωc. For spinning speeds below ωc theory shows that the director makes an increasing angle with the field until at ωc this angle is 45°. Above ωc the director should rotate with an angular velocity slightly less than that of the sample. Observation in both regimes allows ωc to be determined; since it depends on the ratio of the diamagnetic anisotropy to the rotational viscosity coefficient of the nematic, this ratio can be measured. However, an experimental investigation by Eastman et al. [1], suggests that the theoretical relationship between ωc and this ratio may be in error by a factor of about four. We have reanalysed their data in an attempt to check this important claim and have found that there is in fact good agreement between theory and experiment.  相似文献   

8.
The optically active indenyl complexes ((η5-C9H7)Ru(L---L)Cl (where L---L is either (S,S)-1,2-dimethyl-1,2-ethanediylbis(diphenylphosphine) (chiraphos) or (R,R)-1,2-cyclopentanediylbis(diphenylphosphine) (cypenphos)) have been synthesized and spectroscopically characterized and compared with the corresponding cyclopentadienyl complexes. Reaction of the new complexes with 2-e-donors give cationic adducts in which the pentahaptocoordination of the indenyl ligand is maintained. The crystal structures of (S,S)-(η5-C9H7)Ru{Ph2PCH(CH3)CH(CH3)PPh2}Cl (1) and (S,S)-η5-C5H5Ru{Ph2PCH(CH3)CH(CH3)PPh2}Cl (3) have been determined.  相似文献   

9.
Abstract

We have described a theory for U, the potential of mean torque of rigid solutes at infinite dilution in a uniaxial liquid crystal phase; this may be used to calculate (S xx - S yy) and S zz, the principal elements of the Saupe ordering matrix. In its simplest form U(ω) contains only second-rank terms and the dependence of the biaxiality (S xx - S yy) is determined by ω, a parameter which describes the departure of the potential of mean torque from cylindrical symmetry, and is predicted to be temperature independent. If dispersion forces are responsible for the magnitude of the orientational order parameter then ω should be independent of the solvent and depend only on the anisotropy in the electric polarizability of the solute. Indeed, this independence should result for any pair potential which can be factorized into a product of solute and solvent properties. These predictions are tested here by determining values of S zz and (S xx - S yy) for anthracene-d 10 as a solute in several liquid crystal solvents, from the quadrupolar splittings obtained from the deuteron N.M.R. spectra. It is found that ω has a strong dependence on the nature of the solvent, which demonstrates that the solute ordering cannot be determined primarily by dispersion forces, or by a factorizable potential. There is also a weaker temperature dependence of λ observed for each binary mixture, and we show how this might be caused by a dependence of ω on solvent ordering, or by the inclusion of a fourth-rank term in U(ω).  相似文献   

10.
Molecular structures of (triphenylphosphine) [1,1′-bis-(methylthio)ferrocene-S,S′,Fe]Pt(BF4)2 (1), (1,5,9-trithia[9]ferrocenophane-S,S′,S″,Fe)Pd(BF4)2 (2), and (acetonitrile)(1,4,7-trithia[7]ferrocenophane-S,S′,S″,Fe)Pd(BF4)2 (3) were determined by X-ray analyses. The Pt in 1 and the Pd atom in 2 have a somewhat distorted square-planar geometry including the Fe atom of the ferrocene moiety, while the Pd atom in 3 is coordinated by one equivalent of acetonitrile and takes a distorted tetragonal-pyramidal geometry. The distances of the Fe---M bond (M = Pd, Pt) in 1–3 are 2.851(2), 2.827(2), and 3.0962(8) Å, respectively. Cyclic voltammetry of 1–3 gave no reversible wave, but afforded some information supporting the presence of a dative bond.  相似文献   

11.
Measurements of the rotational viscosity γ1 and the density are presented for a mixture of 4'-methoxybenzylidenebutylaniline (MBBA) and its ethoxy homologue EBBA and a mixture of cyclohexylphenylnitriles (ZLI 2413 from Merck AG) as a function of temperature and pressure. A new set-up for the measurement of densities under pressures of up to 3kbar is described. It is shown that the pressure dependence of the kinematic rotational viscosity γ1/ρ and the temperature dependence of γ1 under isobaric and isochoric conditions have common features with that of the shear viscosity of isotropic liquids. Furthermore, it is found that the curves γ1 = f(1/T) for constant p and γ1 = g(ρ) for constant T can be shifted one onto another by an appropriate shift of the scale of the independent variable.  相似文献   

12.
A Doppler-based velocity selection technique has been used to measure the relative velocity dependence of the cross sections σji,Δr) for rotationally inelastic collisions from level ji to ji + Δν1 = 8,22,42) in 7Li*2 A 1Σ+u)—Xe. The σjν±2r) are strongly attenuated at a smaller νr by “torque averaging” due to molecular rotation; in contrast, for large |Δ|, σj = νrn (1 n 2). An empirical intermolecular potential which reproduces these types of behavior for 3-D classical trajectories is exhibited.  相似文献   

13.
The chiral ligands, 4,4′-bis{(1S,2R,4S)-(−)-bornyloxy}-2,2′-bipyridine, (1S,2R,4S)-1, and 4,4′-bis{(1R,2S,4R)-(+)-bornyloxy}-2,2′-bipyridine, (1R,2S,4R)-1, have been prepared and characterized by spectroscopic techniques and, for (1S,2R,4S)-1, by single crystal X-ray diffraction. Despite the use of enantiomerically pure ligands, the formation of the complexes [Fe((1S,2R,4S)-1)3]2+, [Ru((1S,2R,4S)-1)3]2+, [Ru((1S,2R,4S)-1)(bpy)2]2+ and [Ru((1R,2S,4R)-1)(bpy)2]2+ proceeds without preference for either the Δ or Λ-diastereoisomers.  相似文献   

14.
The profiles of the parallel and perpendicular bands ν6 and ν10 of allene in CS2 and their temperature dependence have been studied. These profiles are sensibly lorentzian and consequently allow us to derive the components D and D of the rotational diffusion tensor and the corresponding activation energies; the accuracy of the results is discussed. In particular, it is not possible to determine unambiguously whether the activation energies U| and U are equal or different but, whichever the case, the ratio D/D is of the order of 10. An application test of the extended J rotational diffusion model shows that the results concurrently obtained for the ν6 and ν10 bands are incompatible with this model and we come to the conclusion that the rotation of allene about its various axes is anisotropic.  相似文献   

15.
The structure of 1-chloronaphthalene, C10H7–Cl, at 293 K was investigated using the X-ray diffraction method. Monochromatic radiation MoK (λ=0.71069 Å) enabled the determination of the scattered radiation intensity between S0=4πsin 0/λ=0.430 Å−1 and Smax=14.311 Å−1. The interpretation of the results was carried out using the reduction method of Blum and Narten. Experimental distributions of X-ray scattered intensity were compared with theoretical results predicted for a proposed model of 1-chloronaphthalene molecule. The electron-density radial-distribution function was calculated and some intra- and intermolecular distances in liquid 1-chloronaphthalene were determined. X-ray structural analysis was applied to determine the packing coefficient of 1-chloronaphthalene molecules.  相似文献   

16.
The use of a mono-pivalate mono-acrylate bis-ester of (+)-1S,5S,6S-spiro[4.4]nonane-1,6-diol in an asymmetric Diels-Alder reaction with cyclopentadiene (2 equiv. BCl3, −85°C, CH2Cl2) provided the expected endo bicyclo adduct in >97% de. Iodolactonization of the bicyclo adduct provided the (+)-lactone (5) with a 1S,4S,6S,8R,9S configuration (97% ee). The de's obtained from using various types and amounts of Lewis acids, and both chiral and racemic bis-esters in the Diels-Alder reaction with cyclopentadiene are also reported.  相似文献   

17.
18.
Fluorescent liquid crystalline side chain polymers were synthesized by copolymerization of a ferroelectric monomer and 5 per cent of various blue fluorescent naphthalic imide dye comonomers. Those copolymers were characterized by DSC, X-ray, GPC and optical microscopy. In favourable cases, fast switching fluorescent ferroelectric polymers resulted, exhibiting high tilt angles (up to ∼ 34°) and spontaneous polarization values (up to ∼ 115 nC cm-2) in the S*c phase. One fluorescent copolymer shows orthogonal smectic phases exclusively due to the structure of the incorporated fluorescent comonomer. In this case a strong electroclinic effect and high induced tilt angles (12° 10 V μm-1) have been observed in the Sa phase. Order parameters, S, of the dye moieties up to 0.64 were measured in the room temperature Sb phase for the copolymers  相似文献   

19.
We report measurements of the dynamics of the magnetic Frederiks transition in nematics consisting of disc-like molecules. In this paper the results are presented for three 2, 3, 6, 7, 10, 11-hexakis(p-alkoxybenzoyloxy)triphenylenes, which exhibit a normal nematic phase, and for three 2, 3, 7, 8, 12, 13-hexa(alkanoyloxy)truxenes, which exhibit an inverted nematic phase. We find that the thermal dependence of a bend viscosity coefficient (γ*1) can be accurately described by the expression, γ*1S2 exp (Ea/kT). The absolute value of γ*1 is found to be higher (by a factor of 10-100) than is commonly encountered in nematics consisting of rod-like molecules.  相似文献   

20.
Five novel ruthenium complexes, RuCl2(MOTPP)2[(S,S)-DPEN] [MOTPP = tris(4-methoxyphenyl)phosphine] (1), RuCl2(TFTPP)2[(S,S)-DPEN] [TFTPP = tris(4-trifluoromethylphenyl)phosphine] (2), RuCl2(PPh3)2[(S,S)-DPEN] (3), RuCl2(BDPX)[(S,S)-DPEN] [BDPX = 1,2-bis(diphenylphosphinomethyl)benzene] (4), RuCl2(BISBI)[(S,S)-DPEN][BISBI = 2,2′-bis((diphenylphosphino)methyl)-1,1′-biphenyl]] (5) were synthesized and used for the hydrogenation of aromatic ketones. The complexes showed high catalytic activities, especially that the catalytic activity of complex 5 containing the diphosphine with large bite angle and complex 1 containing triarylphosphine with electron-donating group were higher than the other three complexes. The enantioselectivities of products were almost not influenced by the electron factors of phosphine.  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