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 875 毫秒
1.
A method for the measurement of the rotational viscosity, γ1 of nematic liquid crystals under high pressure is described. First measurements for the liquid crystals 4'-methoxybenzylidene-4-n-butylaniline, a broad range nematic mixture of substituted cyclohexyl-phenyls and a re-entrant nematic mixture are presented.  相似文献   

2.
The crystal structure of the title compound has been determined. Crystals of [HgC12{μ-S(CH2)3NH(CH3)2}] are monoclinic, space group P21/n, with a = 10.136(2), b = 6.519(1), c = 15.940(6) Å and β = 97.20(3)°. The structure consists of (---Hg---S---)n helicoidal chains linked by hydrogen bonding, which give rise to chemically unconnected layers along the ( 02) planes. Each mercury is tetrahedrally coordinated to two terminal chlorine atoms and two bridging sulphur atoms. Assignments of νas(SHgS) and νs(SHgS) for this complex and its isostructural bromine analogue, and of νs(C1HgC1) and νs(BrHgBr) from IR and Raman spectroscopy are reported. Comparison of Hg---S frequencies with those reported for closely related compounds as well as correlation with Hg---S bond distances are made.  相似文献   

3.
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   

4.
K. Czupry&#x  ski  J. Janik  J. K. Mo   cicki 《Liquid crystals》1993,14(5):1371-1375
The phase diagram of a two component system composed of two smectic compounds: 4-octyloxy-4'-cyanobiphenyl (8OCB) and 4-isothiocyanatophenyl 4-butylbenzoate (4TPB) was investigated. Three Miesowicz viscosity coefficients η1, η2, η3 and the refractive indices at different temperatures as well as the enthalpies of the phase transitions were measured. It was stated, that the properties of the induced nematic phase, for example, the nematic phase existing between two smectic regions, are the same as the ones observed in the case of low viscosity nematic mesogens.  相似文献   

5.
In this paper, we report measurements of the viscoelastic properties of nematic liquid crystals which exhibit a glass transition in the nematic phase. We have studied the Freedericksz transition in planar cells with a magneto-optical method. K1 was determined from the critical field, and the rotational viscosity, γ1, from the response time for the director orientation by the external field. We found a temperature dependence of γ1 of the Vogel type, with absolute values ranging over several orders of magnitude, and K1 values similar to those of conventional thermotropic low molar mass nematics.  相似文献   

6.
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   

7.
The light scattering technique was used to investigate the viscoelastic parameters characterizing director twist distortions in miscible nematic mixtures of 5CB (pentacyanobiphenyl) with two side chain liquid crystal polymers and a main chain liquid crystal polymer. By applying an AC electric field to homeotropically-aligned nematic monodomains of the mixtures, the field-dependent scattering intensities and director orientation fluctuation relaxation rates yield, respectively, the twist elastic constant K22 and viscosity coefficient γ1. The results directly demonstrate that the addition of liquid crystal polymers causes substantial decreases of the relaxation rates for dynamic light scattering from the twist mode and these changes are due to small decreases in K22 coupled with large increases in γ1. The decrements in K22 are comparable for both side chain and main chain liquid crystal polymers. The relative increase in the twist viscosity for the side chain liquid crystal polymers is much smaller than those of main chain polymers. A theoretical model is used to qualitatively interpret the difference between the viscous behaviour of the twist mode for both side chain and main chain liquid crystal polymers in a nematic solvent.  相似文献   

8.
The frequency dependence of the dielectric biaxiality of surface stabilized ferroelectric liquid crystals (SSFLCs) was studied. The principal values of the dielectric tensor ε1, ε2 and ε3 were measured by the MOM (molecular orientational model) method. Three dielectric permittivities were measured for each of two samples. These were the permittivity of the homeotropic cell and the permittivity of the planar homogeneous cell with and without the DC bias. Then the dielectric tensor components were calculated based on the molecular orientational models. We present the theory and experimental procedure of the MOM method. Measurements have been performed on Merck FLC compound SCE-8. The following novel dielectric behaviour was observed, as the DC bias voltage was increased the dielectric permittivity of the planar homogeneous cell decreased at the low frequencies (∼ 1 kHz) while increased at the high frequencies (10kHz ∼). The sign of the dielectric biaxiality ∂εε (= ε2 - ε1) inverted around 1 kHz, being negative at low frequencies and positive at high frequencies. The roles of the biaxiality on the dielectric behaviour of SSFLC cells are discussed.  相似文献   

9.
The compound [RU332- -ampy)(μ3η12-PhC=CHPh)(CO)6(PPh3)2] (1) (ampy = 2-amino-6-methylpyridinate) has been prepared by reaction of [RU3(η-H)(μ32- ampy) (μ,η12-PhC=CHPh)(CO)7(PPh3)] with triphenylphosphine at room temperature. However, the reaction of [RU3(μ-H)(μ3, η2 -ampy)(CO)7(PPh3)2] with diphenylacetylene requires a higher temperature (110°C) and does not give complex 1 but the phenyl derivative [RU332-ampy)(μ,η 12 -PhC=CHPh)(μ,-PPh2)(Ph)(CO)5(PPh3)] (2). The thermolysis of complex 1 (110°C) also gives complex 2 quantitatively. Both 1 and 2 have been characterized by0 X-ray diffraction methods. Complex 1 is a catalyst precursor for the homogeneous hydrogenation of diphenylacetylene to a mixture of cis- and trans -stilbene under mild conditions (80°C, 1 atm. of H2), although progressive deactivation of the catalytic species is observed. The dihydride [RU3(μ-H)232-ampy)(μ,η12- PhC=CHPh)(CO)5(PPh3)2] (3), which has been characterized spectroscopically, is an intermediate in the catalytic hydrogenation reaction.  相似文献   

10.
11.
Reactions of Co33-CBr)(μ-dppm)(CO)7 with {Au[P(tol)3]}2{μ-(CC)n} (n=2–4) have given {Co3(μ-dppm)(CO)7}{μ33-C(CC)nC} [n=2 (1), 3 (2), 4 (3)] containing carbon chains capped by the cobalt clusters. Tetracyanoethene reacts with 2 to give {Co3(μ-dppm)(CO)7}233-C(CC)2C[=C(CN)2]C[=C(CN)2]C} (4). X-ray structural characterisation of 1, 3 and 4 are reported, that for 3 being the first of a cluster-capped C10 chain.  相似文献   

12.
Reaction of the activated mixture of Re2(CO)10, Me3NO and MeOH with a 1:1 mixture of rac (d/l)- and meso-1,1,4,7,10,10-hexaphenyl-1,4,7,10-tetraphosphadecane (hptpd) yields a mixture of (d/l)- and meso-[{Re2(μ-OMe)2(CO)6}2(μ,μ′-hptpd)] 1. The diastereomers can be easily separated by selective dissolution of d/l-1 in benzene, and give clearly distinguishable 1H- and 31P-NMR spectra. The fluxional behavior of d/l-1 in solution has been studied by variable-temperature 1H- and 31P-{1H}-NMR spectroscopy. The crystal structures of both d/l- and meso-1 have been determined. Both molecules consist of two {Re2(μ-OMe)2(CO)6} moieties which are bridged by the two P---CH2---CH2---P moieties of the hptpd ligand. Whilst the molecules of meso-1 possess crystallographic i-symmetry, those of d/l-1 do not have any crystallographic symmetry. These diastereomers therefore give clearly distinguishable Raman spectra in the solid state. Reaction of tris[2-(diphenylphosphino)ethyl]phosphine (tdppep) with the activated mixture affords the complex [{Re2(μ-OMe)2(CO)6}(μ,η2-tdppep)] 2, and the analogous reaction involving bis[2-diphenylphospinoethyl)phenylphosphine (triphos) gives [{Re2(μ-OMe)2(CO)6}(μ,μ′,η3-triphos){Re2(CO)9}] 3 and [{Re2(μ-OMe)2(CO)6}(μ,η2-triphos)] 4.  相似文献   

13.
A detailed in situ 13C and 1H NMR spectroscopic characterization of the following families of alkylperoxo complexes of titanium is presented: Ti(η2-OOtBu)n(OiPr)4−n, where n = 1–4; binuclear complexes [(iPrO)3Ti(μ-OiPr)2Ti(OiPr)22-OOtBu)] and [(η2-OOtBu)(iPrO)2Ti(μ-OiPr)2Ti(OiPr)22-OOtBu)]; complexes with β-diketonato ligands: Ti(LL)2(OEt)(η2-OOtBu), Ti(LL)2(OiPr)(η2-OOtBu), Ti(LL)22-OOtBu)2, Ti(LL)2(OtBu)(η1-OOtBu), where HLL = acetylacetone, dipivaloylmethane. These alkylperoxo complexes could not be isolated due to their instability and were studied in situ at low temperatures. Whereas the side-on (η2) coordination mode of tert-butylperoxo ligand is generally preferable, the end-on (η1) coordination caused by spatial hindrance from surrounding bulky ligands is found in two cases. The quantitative data on the reactivity of alkylperoxo complexes found towards sulfides and alkenes were obtained. The system TiO(acac)2/tBuOOH in C6H6 was reinvestigated using 13C and 1H NMR spectroscopy. The structure of the complex Ti(acac)2{CH3C(O)(OOtBu)COO} actually formed in this system was elucidated. Four types of titanium(IV) alkylperoxo complexes were detected in the Sharpless–Katsuki catalytic system using 13C NMR spectroscopy.  相似文献   

14.
Reaction of [Ru3(CO)12 with (CF3)2P---P(CF3)2 in p-xylene at 140°C yielded the compounds [Ru4(CO)13{μ-P(CF3)2}2] (1), [Ru4(CO)14{μ-P(CF3)2}2] (2) and [Ru4(CO)11{μ-P(CF3)2}4] (3). Reaction with [(μ-H)4Ru4(CO)12] under similar conditions yielded [(μ-H)3Ru4(CO)12{μ-P(CF3)2}] (4). All four compounds have been characterised by X-ray crystallography. The fluxional behaviour of the hydrides in 4 has also been studied by variable-temperature NMR spectroscopy. Compounds 1, 2 and 4 were also obtained from the reactions of Ru3(CO)12 with (CF3)2PH in dichloromethane at 80°C.  相似文献   

15.
The reactions of the diruthenium carbonyl complexes [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]X (X=BF4 (1a) or PF6 (1b)) with neutral or anionic bidentate ligands (L,L) afford a series of the diruthenium bridging carbonyl complexes [Ru2(μ-dppm)2(μ-CO)22-(L,L))2]Xn ((L,L)=acetate (O2CMe), 2,2′-bipyridine (bpy), acetylacetonate (acac), 8-quinolinolate (quin); n=0, 1, 2). Apparently with coordination of the bidentate ligands, the bound acetate ligand of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ either migrates within the same complex or into a different one, or is simply replaced. The reaction of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ (1) with 2,2′-bipyridine produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)2] (2), [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-bpy)]+ (3), and [Ru2(μ-dppm)2(μ-CO)22-bpy)2]2+ (4). Alternatively compound 2 can be prepared from the reaction of 1a with MeCO2H–Et3N, while compound 4 can be obtained from the reaction of 3 with bpy. The reaction of 1b with acetylacetone–Et3N produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-acac)] (5) and [Ru2(μ-dppm)2(μ-CO)22-acac)2] (6). Compound 2 can also react with acetylacetone–Et3N to produce 6. Surprisingly [Ru2(μ-dppm)2(μ-CO)22-quin)2] (7) was obtained stereospecifically as the only one product from the reaction of 1b with 8-quinolinol–Et3N. The structure of 7 has been established by X-ray crystallography and found to adopt a cis geometry. Further, the stereospecific reaction is probably caused by the second-sphere π–π face-to-face stacking interactions between the phenyl rings of dppm and the electron-deficient six-membered ring moiety of the bound quinolinate (i.e. the N-included six-membered ring) in 7. The presence of such interactions is indeed supported by an observed charge-transfer band in a UV–vis spectrum.  相似文献   

16.
The acid–base chemistry of some ruthenium ethyne-1,2-diyl complexes, [{Ru(CO)2(η-C5H4R)}22-CC)] (R=H, Me) has been investigated. Initial protonation of [{Ru(CO)2{η-C5H4R}}22-CC)] gave the unexpected complex cation, crystallised as the BF4 salt, [{Ru(CO)2(η-C5H4R}}33-CC)][BF4] (R=Me structurally characterised). This synthesis proved to be unreliable but subsequent, careful protonation experiments gave excellent yields of the protonated ethyne-1,2-diyl complexes, [{Ru(CO)2{η-C5H4R)}2212-CCH)](BF4) (R=Me structurally characterised) which could be deprotonated in high yield to return the starting ethyne-1,2-diyl complexes.  相似文献   

17.
The chemistry of the di-μ-methylene-bis(pentamethylcyclopentadienyl-rhodium) complexes is reviewed. The complex [{(η5-C5Me5)RhCl2}2] (1a) reacted with MeLi to give, after oxidative work-up, blood-red cis-[{(η5-C5Me5)Rh(μ-CH2)}2(Me)2], 2. This has the two rhodiums in the +4 oxidation state (d5), and linked by a metal-metal bond (2.620 Å). Trans-2 was formed on isomerisation of cis-2 in the presence of Lewis acids, or by direct reaction of 1a with Al2Me6, followed by dehydrogenation with acetone. The Rh-methyls in [{(η5-C5Me5)Rh(μ-CH2)}2(Me)2] were readily replaced under acidic conditions (HX) to give [{(η5-C5Me5)Rh(μ-CH2)}2(X)2] (X = Cl, Br or I); these latter complexes reacted with a variety of RMgX to give [{(η5-C5Me5)Rh(μ-CH2)}2(R)2] (R = alkyl, Ph, vinyl, etc.). Trans-2 also reacted with HBF4 in the presence of L to give first [{(η5-C5Me5)Rh(μ-CH2)}2(Me)(L)]+ and then [{(η5-C5Me5)Rh(μ-CH2)}2(L)2]2+ (L = MeCN, CO, etc.). The {(η5-C5Me5)Rh(μ-CH2)}2 core is rather kinetically inert and also forms a variety of complexes with oxy-ligands, both cis-, e.g. [{(η5-C5Me5)Rh(μ-CH2)}2(μ-OAc)]+ and trans-, such as [(η5-C5Me5)Rh(μ-CH2)}2(H2O)2]2+. The complexes [{(η5-C5Me5)Rh(μ-CH2)}2(R)L]+ (R = Me or aryl) in the presence of CO, or [{(η5-C4Me5)Rh(μ-CH2)}2(R)2] (R = Me, Ph or CO2Me) in the presence of mild oxidants, readily yield the C---C---C coupled products RCH=CH2. The mechanisms of these couplings have been elucidated by detailed labelling studies: they are more complex than expected, but allow direct analogies to be drawn to C---C couplints that occur during Fischer-Tropsch reactions on rhodium surfaces.  相似文献   

18.
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   

19.
Reaction of the ruthenium(IV) chloro-bridged dimer [{Ru(η3 : η3-C10H16)Cl(μ-Cl)}2], 1, with ethanethiol (EtSH) in CH2Cl2 gives the bridged-cleaved adduct [Ru(η3 : η3-C10H16)Cl2(SHEt)], 2. Stirring of two molar equivalents of 2 in methanol with one equivalent of 1 gives the binuclear, mixed chloro/thiolato bridged compound [{Ru(η3 : η3-C10H16)Cl} 2(μ-SEt)], 3. The related doubly thiolato bridged complex [{Ru(η3 : η3-C10H10)Cl(μ-SEt)}2], 4, is formed by treatment of 1 with an excess of EtSH, or by prolonged stirring of 2 alone in methanol. Compounds 2–4 have been studied by cyclic voitammetry. Compound 2 undergoes only irreversible oxidation, whereas in the case of both 3 and 4 the observation of significant return waves is consistent with a greater stability of the primary redox products.  相似文献   

20.
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   

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