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 125 毫秒
1.
Polycrystalline alunite‐d6 KAl3(OD)6(SO4)2, prepared by hydrothermal reaction of Al2(SO4)3, K2SO4 and D2SO4, was studied by neutron powder diffraction performed on the diffractometer E2 (HMI‐BENSC, Berlin). Rietveld refinement of the data set for T = 2 K yielded the crystallographic data: space group R3m, Z = 3, trigonal setting, a = 694.3(1) pm, c = 1722.7(2) pm, N(I/σ(I) > 1) = 172, N(Var.) = 19, Rp = 0.036, wRp = 0.046, RB(I/σ(I) > 1) = 0.020. The deuterium nuclei could be located precisely. Three equivalent O–D bonds with nuclear distances r(O(4)–D) = 96.6(3) pm directed to each of the terminal oxygen atoms of the SO4 groups are found. Partial substitution of K+ by D3O+ was also considered in the refinement procedure. In good agreement with results of other methods a site occupation fraction n(D3O+) = 0.0104 was obtained.  相似文献   

2.
The oxides A(Ti0.5Te1.5)O6 (A = K, Rb, Cs, Tl), A(Ti0.5W1.5)O6 (A = Rb, Cs, Tl), and Cs(B0.5W1.5)O6 (B = Zr, Hf) have been obtained as polycrystalline powders giving X-ray diffraction patterns characteristic of defect cubic pyrochlores, space group (No. 227), Z = 8. The best discrepancy R factors, from 0.0265 for Rb(Ti0.5Te1.5)O6 to 0.0554 for Cs(Zr0.5W1.5)O6, were obtained for the B cations randomly distributed at 16(d), A ions at one quarter of 32(e), and oxygen atoms at 48(f) positions. A linear relationship is observed between the a unit cell parameters and the ionic radii of the A cations, as well as the average ionic radii of the B atoms. The results of electrical resistivity measurements for A(Ti0.5Te1.5)O6 (A = K, Rb, Cs, Tl) are given.  相似文献   

3.
The frontier orbital interactions of electron pushing and drawing substituents with ferrocenyl group were analyzed based on the electrochemical,UV visible spectral and spectroelectrochemical results of five ferrocene derivatives,R-Fc-A1(PⅠ),A1-Fc-A1(PⅡ),D-Fc-R (PⅢ),D-Kc-A1(PIV) and D-Fc-A2(PV)(R,CH2OH;A1 CHO;A2,CH=C(CN)2 and D,(C18H37)2N-C6H4-CH=CH) It was found that there are strong interactions of the LUMO (πA) of electron drawing substituents with le2g(dxy,dx2 y2)and e2u of the ferroeenyl group because the energy levels of πA and e2g,C2U of (Cp )2 are close,which lower not only the energy levels of bonded orbits,πA+ and dx2-y2+[πA] of PⅠ,PⅡ,PⅣ and PⅤobviously,but also those of their non-bonded orbu dxy For PⅢ,PⅣ and PⅤ,there are strong interactions of HOMO(πD) of the electron pushing substituent with le of the ferrocenyl group because the levels of πD and e of (Cp)2 are close,which result in the formation of anti-bonded orbit,πD- and bonded orbit  相似文献   

4.
A kinetic study of the very low-pressure pyrolysis of ethylbenzene (I), 2-phenylethylamine (II), and N,N-dimethyl 2-phenylethylamine (III) above 900 K yields the heats of formation of aminomethyl (A) and N,N-dimethylaminomethyl (B) radicals: ΔH?, 300 K(A) = 30.3 and ΔH?, 300 K(B) = 27.5 kcal/mol. The difference of stabilization energies Es, (relative to methyl radicals): Δ = Es(B) ? Es(A) = (2 ± 1) kcal/mol, conforms to similar effects in methyl substituted alkyl and amino free radicals.  相似文献   

5.
Fe2(CO)9 and R2P(S)P(S)R2 (R = Et, n-Pr, n-Bu, Ph) react to form two types of cluster complexes Fe3(CO)93-S)2 (1), Fe2(CO)6(μ-SPR2)2 (2A)–(2D), [2A, R = Et; 2B, R = n-Pr; 2C, R = n-Bu; 2D, R = Ph]. The complexes result from phosphorus–phosphorus bond scission; in the former sulfur abstraction has also occurred. The complexes have been characterized by elemental analyses, FT-IR and 31P-[1H]-NMR spectroscopy and mass spectrometry.  相似文献   

6.
用对称透射法,测试聚酯(PET)取向非晶试样的二维广角X射线散射相干强度I(K,α),将不同方位角α上的I(K,α)非晶重叠峰曲线进行计算分解,获得分子链间原子散射引起的二个峰为A(K=0.126),B(K=0.169),分子链内原子散射引起五个峰分别为C(K=0.304),D(K=0.553),E(K=0.374),F(K=0.465),G(K=0.606),K单位为nm^-1,分析了它们的结构  相似文献   

7.
New chiral diaza-18-crown-6 ether derivatives, 5 and 6 were synthesized from (R)-(-)-2-amino-1-bütanol. These chiral artificial receptors exhibit pronounced chiral recognition toward the enantiomers of l- and d- amino acid derivatives. The highest enantioselectivity was observed in the case of Trp-OMe·HCl (KD/KL=12.5).  相似文献   

8.
Photon correlation spectroscopy is employed to study the slowly relaxing density and anisotropy fluctuations in bulk atactic polystyrene as a function of temperature from 100 to 160°C and pressure from 1 to 1330 bar. The light-scattering relaxation function is well described by the empirical function ?(t) = exp[?(t/τ)β], where for polystyrene β = 0.34. The average relaxation time is determined at each temperature and pressure according to 〈τ〉 = (τ/β)Γ(1/β) where Γ(x) is the gamma function. The data can be described by the empirical relation 〈τ〉 = 〈τ〉0 exp[(A + BP)/R(T ? T0)] where R is the gas constant and T0 is the ideal glass transition temperature. The empirical constant A/R is in good agreement with that determined from the viscosity or the dielectric relaxation data (1934 K). The empirical constant B can be interpreted as the activation volume for the fundamental unit involved in the relaxation and is found to be comparable to one styrene subunit (100 mL/mol). The quantity B appears to be a weak function of temperature. The use of pressure as a tool in the study of light scattering near the glass transition now has been established.  相似文献   

9.
The title compoundsR 2Si(NHR 1) (NHR 2) (B) were prepared according to Equ. (2) and (3) and via formerly unknown diphenyl-chloro-organylaminosilanes (A) as intermediates. The complex course of formation ofB in competition withR 2Si (NHR 1)2 (C) andR 2Si(NHR 2)2 (D) (Scheme 1, Equ. (4)–(7), Table 2) was investigated in detail. Results of thermal rearrangement ofB are given in Table 5. Five novelA and six novelB compounds are confirmed by properties, elemental and structural data (Tables 1, 3 and 4).
Mit Auszügen aus der DissertationS. Klemke, Techn. Univ. Braunschweig 1978.  相似文献   

10.
Linear dichroism (LD) spectra are presented for naphthalene oriented in stretched polyethylene and polypropylene matrices at 77 K and 296 K. From the calculated spectrum LD(λ)/A(λ), where A(λ) is the corrected absorbance spectrum of the sample by unpolarized light, orientational parameters are calculated and component spectra, 235–315 nm, are resolved corresponding to polarization parallel to the long (B3u = x) and the short (B2u = y) axes in the molecular plane (D2h). The orientational parameters indicate different orientational mechanisms in polyethylene and polypropylene, but the resolving procedure yields mainly identical component spectra. It is suggested that the polarization (B3u) predominating in the 245–275 nm region isdue to a B1g vibronic perturbation of the 1B2u state.  相似文献   

11.
SCF closed shell calculations were performed to determine the equilibrium structure and vibrational frequencies of the O4 molecule by means of Payne's method and with the help of the molecule's symmetry coordinates. The equilibrium geometry corresponds to symmetry group D2d with R = 1.505 Å and h = 0.094 Å. The vibrational frequencies are: ν5(E) = 885.5 cm?1, ν3(B1) = 1051.9 cm?1, ν1(A1) = 1018.3 cm?1, ν4(B2) = 880.3 cm?1. The second vibrational coordinate (A1) corresponds to a double-well potential. The first vibrational levels were calculated by a variational method.  相似文献   

12.
In order to evaluate the effectiveness of l-lactate dehydrogenase (LDH) from rabbit muscle as a regenerative catalyst of the biologically important cofactor nicotinamide adenine dinucleotide (NAD), the kinetics over broad concentrations were studied to develop a suitable kinetic rate expression. Despite robust literature describing the intricate complexations, the mammalian rabbit muscle LDH lacks a quantitative kinetic rate expression accounting for simultaneous inhibition parameters, specifically at high pyruvate concentrations. Product inhibition by l-lactate was observed to reduce activity at concentrations greater than 25 mM, while expected substrate inhibition by pyruvate was significant above 4.3 mM concentration. The combined effect of ternary and binary complexes of pyruvate and the coenzymes led to experimental rates as little as a third of expected activity. The convenience of the statistical software package JMP allowed for effective determination of experimental kinetic constants and simplification to a suitable rate expression:
v = \fracVmax( AB )KiaKb + KbA + KaB + AB + \fracPKI - Lac + \fracB2AKI - Pyr + \fracB2QKI - Pyr - NAD v = \frac{{{V_{max}}\left( {AB} \right)}}{{{K_{ia}}{K_b} + {K_b}A + {K_a}B + AB + \frac{P}{{{K_{I - Lac}}}} + \frac{{{B^2}A}}{{{K_{I - Pyr}}}} + \frac{{{B^2}Q}}{{{K_{I - Pyr - NAD}}}}}}  相似文献   

13.
17O-NMR spin-lattice relaxation timesT 1 of D2O molecules were measured at 5–85°C in D2O solutions of alkali metal halides (LiClCsCl, KBr, and KI), DCl, KOD, Ph4PCl, NaPh4B, and tetraalkylammonium bromides (Me4NBrAm4NBr) in the concentration range 0.1–1.4 mol-kg–1 TheB-coefficients of the electrolytes obtained from the concentration dependence of relaxation ratesR 1=1/T1 were divided into the ionicB-coefficients by three methods: (i) the assumption ofB (K+)=B(Cl), (ii) the assumption ofB(Ph4P+)=B(Ph4B), and (iii) the use ofB(Br) obtained from a series ofB(R4NBr). It was found that Methods (ii) and (iii) resulted in an abnormal temperature dependence of theB-coefficients of alkali metal ions and a negative values of rotational correlation times c at lower temperatures for hydroxide and halide ions. These results suggest that the methods based on the van der Waals volume are not adequate for the ionic separation of NMRB-coefficients. From the analysis using the assumption ofB(K+)=B(Cl), it was found that D3O+, OD, and Me4N+ ions are the intermediates between structure makers and breakers, and that the hydrophobicity of phenyl groups is weaker than that of alkyl groups due to the interactions between water molecules and -electrons in phenyl groups.  相似文献   

14.
The hydrogen bonding abilities of iV-methylimidazolidin-2-one and -2-selone and thiazolidin-2-one and -2-selenone have been studied by ir spectroscopy at 25° in carbon tetrachloride solutions, using dimethyl sulphoxide and 4-chlorophenol as proton acceptor (KA) and proton donor (KB), respectively. The results are compared with those previously reported for N-methylimidazolidine-2-thione and thiazolidine-2-thione. The KA values increase in each series in the order O < S < Se and Kg in the reverse order. The and KB values are discussed in terms of the substituent in ring. The self-association constants (KD) are dependent on both KA and KB, although KA seems to be much more important.  相似文献   

15.
It is shown under very general conditions that the intermediate scattering function for the generalized Rouse—Zimm model always takes the simple form G(K, t) α exp[?K2(kBT/f)t], when the scattering vector K becomes sufficiently large. (Here kB is Boltzmann's constant, T is the absolute temperature and f is the individual bead friction factor.) A microscopic formulation for the bulk modulus and friction factor density of a gel network is incorporated into the viscoelastic continuum model of Tanaka et al. The resulting expression for the apparent long-wavelength diffusion coefficient of the gel is DG = (kBT/f)2(1 - 2/Φ), where Φ is the network functionality.  相似文献   

16.
Crystals of mixed alkali neodymium orthoborates, K9Li3Nd3(BO3)7 and A2LiNd(BO3)2 (A = Rb, Cs) were obtained by spontaneous crystallization. K9Li3Nd3(BO3)7 crystallizes in space group P2/c with cell parameters of a = 11.4524(7) Å, b = 10.1266(6) Å, c = 12.3116 (10) Å, β = 122.0090(10)°. In the structure, NdO8 polyhedra share corners and connect with planer BO3 groups to form infinite [Nd3B3O21]n chains. These chains are linked by additional BO3 groups to produce a double layer of [Nd6B6O38]n blocks in the ac plane with K and Li ions filled into the cavities. A2LiNd(BO3)2 (A = Rb, Cs) crystallizes in space group Pbcm, with cell parameters of a = 7.113(2) Å, b = 9.691(3) Å and c = 10.135(3) Å for Rb2LiNd(BO3)2, and a = 7.2113(3) Å, b = 9.9621(4) Å, and c = 10.3347(4) Å for Cs2LiNd(BO3)2. In the structure, NdO8 polyhedra are corner‐sharing with each other and further interlinked by BO3 groups to comprise the infinite [Nd4B4O24] sheets in the bc plane, with Rb/Cs and Li ions occupying the interlayered space. The compounds show effective near‐IR emission and their associated lifetimes are obtained by fluorescence spectra.  相似文献   

17.
Equilibria of Mo(VI) in acid aqueous solutions with excess of 2,3-dihydroxynaphthalene (DHN)c DHN /c Mo = 2.3–107 (I = 0.6 mol 1–1 (NaClO4), 0.6% v/v ethanol) were studied spectrophotometrically. Formation constants of MoO2R 2 2– (logK 012 = 5.89±0.01) and presumed MoO2(OH)(OH2)R (logK 111 = 7.79±0.01) chelates were evaluated using SQUAD-G program.  相似文献   

18.
0IntroductionNiobateshavemanykindsofstructures.TheseareperovskitetypeABO3(KNbO3),tungstenbronzetypeANb2O6(Ba2NaNb5O15),chainandlamellartypes犤1~11犦.SincethenewcompoundK6CrNb15O42withakindoftunnelstructureinthepotassiumniobatesystemwasfoundinourlaboratoryforthefirsttime犤12犦,wehavesynthesizedaseriesofcompoundswiththesamestructure,forexample,K6FeNb15O42,K6Ni0.67Nb15.33O42,Ba6Cr4Nb12O42andBa6Ni2.67Nb13.33O42etc.T…  相似文献   

19.
The average interatomic distances D in oxygen polyhedra MO n of isostructural oxides were proposed to be estimated using the equation D = Kɛ(R M + R O) or D = Ax 2 + Bx + C, where x = ɛ(R M + R O), ɛ is the ionicity of the M-O bond, R M is the ionic radius of the cation M with account for the coordination in the polyhedron, and R O is the ionic radius of oxygen. Calculations were made for MO oxides having the rock-salt structure; Ln2O3 oxides, where Ln = Ce-Yb; and the MO2 oxides having the rutile and fluorite structures.  相似文献   

20.
Dynamic light scattering experiments have been performed at various concentrations, of pharmaceutical oil-in-water microemulsions consisting of Eutanol G as oil, a blend of a high (Tagat O2) and a low (Poloxamer 331) hydrophilic–lipophilic balance surfactant, and a hydrophilic phase (propylene glycol/water). We probe the dynamics of these microemulsions by dynamic light scattering. In the measured concentration range, two modes of relaxation were observed. The faster decaying mode is ascribed classically to the collective diffusion D c (total droplet number density fluctuation). We show that the slow mode is also diffusive and suggest that its possible origin is the relaxation of polydispersity fluctuations. The diffusion coefficient associated with this mode is then the self-diffusion D s of the droplets. It was found that D c and D s had opposite volume fractions of oil plus surfactants (ϕ) dependence and a common limiting value D 0 for ϕ=0. Average hydrodynamic radius (R h=10.5 nm) of droplets was calculated from D 0. R h is supposed to compose the inner core, a surfactant film including possible solvent molecules, which migrate with the droplet. The concentration dependence of diffusion coefficients reflects the effect of hard sphere and the supplementary repulsive interactions which arises due to loss of entropy, when absorbed chains of surfactant intermingle on the close approach of the two droplets. This mechanism could also explain the observed stability of our systems. The estimated extent of polydispersity is 0.22 from the amplitude of slower decaying mode. The polydispersity in microemulsion systems is dynamic in origin. Results indicate that the time scale for local polydispersity fluctuations is at least three orders of magnitude longer than the estimated time between droplet collisions.  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