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 31 毫秒
1.
Relative rates of solvolysis of some N-triorganosilylanilines in mixtures of ethanol and aqueous potassium hydroxide have been determined, with results as follows. (i) For XC6H4NHSiEt3 compounds in MeOH (5 vol) + aq. alkali (2 vol) at 50°: (X =) H, 1.0;p-Me, 0.80;p-OMe, 0.83;m-Me, 0.90; o-Me, 0.87; p-SMe, 1.90; p-F, 1.7; p-Cl, 2.8; o-Cl, 14; m-Cl, 4.2; m-NO2, 18; p-CN, ca. 43; p-NO2, ca. 120. (ii) For PhNHSi(C6H4Y)3 compounds in MeOH (10 vol) + aq. alkali (1 vol) at 50°: (Y =) H, 1.0; p-OMe, 0.12; p-Cl, ca. 32; m-Cl, ca. 84. (iii) For PhNHSiR3 compounds in MeOH (5 vol) + aq. alkali (2 vol) : (R3 =) Et3, 1.0; Et2Me, 18 (at 30°); Me2-i-Pr, 8 (at 30°);Me2-t-Bu, 0.012 (at 50°);i-Pr3, 0.006 (at 50°). In series (i) the relative rates correlate with σ, or where appropriate σ?-constants, with a ? value of 1.6. It is suggested that in the transition state of the rate-determining step the OSi bond is fully formed, or almost so, the SiN bond approximately 20—50% broken, and the bond between the nitrogen atom and a proton from the solvent ca. 10—30% formed.  相似文献   

2.
A stereochemical study of a series of alkyl glucosyl-β-(1→6)-thioglucosides and their S-oxides by means of nuclear magnetic resonance and circular dichroism revealed that the populations around the thioglucosidic bond (ring I) as well as those of the interglycosidic linkage ω depend on the aglycone, the solvent and, in the S-oxides, on the absolute configuration of the sulfur atom. The results for the thio-disaccharides showed the strong influence of the solvent polarity on the conformational preferences of the interglycosidic bond. In polar solvents, the magnitudes of the rotamer populations, Pgg and Pgt, remained practically constant through the series, while in non-polar solvents a clear predominance of gt conformation was observed as well as the influence of the aglycone on the conformational equilibrium. The results for both (SS)- and (RS)-alkyl thiogentiobiosyl S-oxide series showed a clear predominance of the gt rotamer, Pgt always having a higher magnitude in the latter series than in the former. Both series exhibited linear correlations between interglycosidic Pgg and Pgt and Taft’s steric parameter (ES) for the alkyl group attached to the sulfinyl group, especially in non-polar solvents. The stereochemical study around the C1–S bond established that the flexibility around this linkage depends on aglycone size, solvent polarity, and the absolute configuration of the sulfur, derivatives of the SS series showing higher flexibility in both polar and non-polar media. The conformational properties of these compounds in solution are explained in terms of stereoelectronic, steric, and solvent effects.  相似文献   

3.
Vinyl acetate was polymerized in bulk and in benzene at 50°C using a wide range of concentrations of azobisisobutyronitrile. Values of fk (the efficiency of initiator) and kprt/kikp (the characteristic constant of primary radical termination) were found to be 0.53 and 2.00 × 104 respectively from data for bulk polymerization. In solution polymerization, the initiator exponent is a function of initiator concentration ranging from 0.35 at high concentration to- about 0.65 at low concentration. This result has been explained on the basis of degradative chain transfer to solvent and primary radical termination. The results have been treated according to mathematical formulations already developed; the characteristic constant of degradative chain transfer and the transfer constant of the solvent have been determined. The results have been compared with literature values and discrepancies explained.  相似文献   

4.
The liquid-phase catalytic hydrogenation of acetylene into ethylene in the presence of CO over palladium supported on the graphite-like material Sibunit has been investigated. Carbon monoxide is an effective modifier of the selective hydrogenation process, exerting its effect by competing with acetylene and ethylene for chemisorption sites on the palladium surface. Under the optimum conditions (T = 90°C; N-methylpyrrolidone solvent; feed consisting of 2 vol % C2H2, 90 vol % H2, and He balance), the introduction of 2 vol % CO ensures a high ethylene selectivity of 89.6 ± 1.5% at an acetylene conversion of 95.8 ± 1.3%, with the acetylene converted into hydrooligomers taken into account.  相似文献   

5.
The Fermi doublet V2?V3 + V4 of CH3CN in basic, inert and acidic solvents has been studied by IR and Raman spectroscopy. The values of W, the Fermi coupling coefficient, obtained from IR spectra varies with the nature of the solvent while W evaluated from Raman data remains constant at 12.5 ± 0.5 cm?1. The similar effects of Bronsted and Lewis acids on the band frequencies and intensities is evidence that the CN group complexes with acids via the N atom “n” electron pair and not the π bond.  相似文献   

6.
An equilibrium study on complexation behavior of the inner transition metals, where M = {Y(III), La(III, Ce(III, Pr(III), Nd(III), Sm(III), Gd(III), Dy(III), Th(IV)} and L = piracetam, has been carried out using the Irving–Rossotti titration technique in aqueous media at different temperatures and ionic strengths. The protonation constant and stability constants (log10 β n ) of the resulting inner metal–ligand complexes have been calculated with the Fortran IV program BEST using the method of least squares and considering the BESTFIT model. Thermodynamic parameters were also evaluated, yielding negative ΔG°, ΔH° and positive ΔS° values that indicate complex formation is favorable at ordinary temperatures. Species distribution curves of complexes have been plotted as function of pH using the SPEPLOT FORTRAN IV program to visualize the presence of various species in equilibrium in the pH range 2–6. To understand more about these equilibria, the stability of these complexes was studied at different percentages of solvent (10–30 % v/v) in different aqua-organic solvent mixtures with methanol, ethanol, 1-butanol and dimethylformamide.  相似文献   

7.
《Solid State Sciences》2000,2(3):385-389
The most commonly used equations correlating bond valence and bond length have been critically compared. It has been shown that the Zachariasen equation is more accurate than the Brown–Shannon equation. Doubts already voiced about the universality of the constant B in the Brown–Altermatt equation with a value of 0.37 Å have been hereby confirmed. Moreover, by a method based on the comparison of formal oxidation states and valences of molybdenum in suitable oxides, the parameters relative to the Zachariasen equation have been accurately determined for the MoO bond in MoO6 octahedra. Their values are R1=1.8790 and B=0.3048 Å in the 3–6 v.u. range.  相似文献   

8.
The photolysis of 2,2,4,6-tetramethyl-1,2-dihydroquinoline in binary MeOH-C5H12 and MeOH-MeCN mixtures was studied in dependence on the solvent composition by steady state and pulse photolysis. The photoinduced proton transfer from the N-H bond to the C(3) carbon atom of the heterocycle does not occur up to the methanol concentration of 0.25 (1 vol %) and 2.5 mol l−1 (10 vol %) in the MeOH-C5H12 and MeOH-MeCN mixtures, respectively. The trends in the increase in the relative quantum yield of the proton transfer reaction (Q mix/Q MeOH) and the decay rate constant for 2,2,4,6-tetramethyl-2,3-dihydroquinoline, the product of the proton transfer reaction, in dependence on the composition differ significantly for these binary mixtures. The results are interpreted in terms of peculiarities in the aggregation of methanol and the distribution of the DHQ molecules and transient species in the mixtures.  相似文献   

9.
The kinetics of amide bond cleavage of isatin and N-methylisatin in the presence of N,N-dimethylacetamide (DMA) was followed spectrophotometrically in the range of solvent composition (0–48.53 wt.%) and temperatures (40–70 °C) using piperidine as a nucleophile. The reaction was studied under pseudo-first-order kinetics. The rate of the reaction decreases largely with increasing organic solvent content. The thermodynamic activation parameters were calculated and discussed in terms of solvation of the activated complex. No linearity was observed between log rate constant and the reciprocal dielectric constant for the solvent used suggesting that there is a selective solvation by higher polar solvent (water). Finally, a mechanism for the ring opening for isatin and N-methylisatin was proposed.  相似文献   

10.
The concept of bond order has been extended to Del Re calculations, and the bond orders of tin-chlorine and tin-carbon bonds in Me4-nSnCln (n = 1 to 4) type compounds have been calculated. The tin-chlorine bond order increases progressively from 0.922 in Me3SnCl to 0.977 in SnCl4, and correlates satisfactorily with the experimental tin-chlorine bond distances. The tin-carbon bond order, on the other hand, remains almost constant, in agreement with the constancy of tin-carbon bond distance in the series. The average tin-chlorine, tin-carbon and tin-hydrogen stretching frequencies in similar compounds vary linearly with the calculated bond polarities indicating variation in bond polarity to be the dominating factor. The unusually low values of the tin-carbon stretching frequency for the tin-vinyl bond compared to the tin-methyl bond in Me3ViSn and Et2Vi2Sn can also be explained in terms of larger polarity of the tin-vinyl bond in these compounds.  相似文献   

11.
An ab-initio gaussian lobe function study of the structure and some molecular properties of the HPO molecule is reported. By a partial optimization of the geometrical structure the following equilibrium structure parameters are found: rPO = 1.536 Å; rPH = 1.475 Å; the bond angle is found to be close to 102° with a bending force constant of 0.58 mdyne/Å. The dipole moment, effective charges and quadrupole term components at various bond angles are reported. The wavefunction of the equilibrium geometry is available from the authors upon request.  相似文献   

12.
We present data for a comparison of a liquid-chromatographic method coupled with tandem mass spectrometry (LC-MS/MS) and a high-performance liquid-chromatographic method with column switching and UV spectrophotometric detection. The two methods were developed for determination of naltrexone and 6β-naltrexol in blood serum or plasma aiming to be used for therapeutic drug monitoring to guide the treatment of patients with naltrexone. For the high-performance liquid chromatography (HPLC)/UV detection, online sample cleanup was conducted on Perfect Bond C18 material with 2% (vol/vol) acetonitrile in deionized water. Drugs were separated on a C18 column using 11.5% (vol/vol) acetonitrile and 0.4% (vol/vol) N,N,N,N-tetramethylethylenediamine within 20 min. LC-MS/MS used naltrexone-d 3 and 6β-naltrexol-d 4 as internal standards. After protein precipitation, the chromatographic separation was performed on a C18 column by applying a methanol gradient (5–100%, vol/vol) with 0.1% formic acid over 9.5 min. The HPLC/UV method was found to be linear for concentrations ranging from 2 to 100 ng/ml, with a regression correlation coefficient of r 2?>?0.998 for naltrexone and 6β-naltrexol. The limit of quantification was 2 ng/ml for naltrexone and 6β-naltrexol. For the LC-MS/MS method the calibration curves were linear (r²?>?0.999) from 0.5 to 200 ng/ml for both substances, and the limit of quantification was 0.5 ng/ml. The concentrations measured by the two methods correlated significantly for both substances (r²?>?0.967; p?<?0.001). Both methods could be used for therapeutic drug monitoring. The HPLC/UV method was advantageous regarding automatization and costs, whereas LC-MS/MS was superior with regard to sensitivity.  相似文献   

13.
The preferential solvation approach and the dielectric enrichment model have been applied to explain the solvatochromic behavior of o-, m- and p-nitroaniline (oNA, mNA and pNA) in several binary solvent mixtures. Cyclohexane was used as the “inert” nonpolar cosolvent in every mixture. The other solvents were chosen trying to vary their polarity as much as possible as well as their hydrogen bond donor or acceptor capabilities. Preferential solvation is detected in every solvent mixture studied. These global interactions were quantified by calculating the preferential solvation constant, K. Also, by using the previously developed model, we calculated for each pair of solvent mixtures a theoretical curve and the corresponding KD due to dielectric enrichment. Non hydrogen bond acceptor solvents (β=0), give values of K quite similar to those of KD, indicating that the preferential interaction is practically dielectric in nature. When the interacting solvent is a hydrogen bond acceptor, the values of K are higher than KD according to the acidity of the H in the amino group in the solutes. The values of K as well as of KD for any solvent mixtures in general follow the order pNA > mNA > oNA as expected, considering the values of μg and μgex. Studies in pure solvent support previous conclusions.  相似文献   

14.
The electrical conductances of tris-(ethylenediamine)chromium chloride, ([Cr(en)3]Cl3; en = ethylenediamine) were measured as functions of temperature and concentration in 20–80 wt% dimethyl sulfoxide (DMSO)water mixtures. Conductance data were analyzed by the Kraus–Bray and Shedlovsky models. Also, density and viscosity values for 20–80 wt% DMSO–water mixtures have been determined experimentally at temperatures varying from (288.15 to 318.15) K. The limiting molar conductance, Λ 0, and the association constant, K A, for ([Cr(en)3]Cl3 were computed by using Shedlovsky’s equation. The Λ 0 values decrease with increase in the percentage of DMSO in the mixed solvent. The K A values increase with increasing temperature and increasing percentage of DMSO in the mixed solvent at all temperatures. This may be attributed to the increase in association constant with the decrease in the relative permittivity of the mixed solvent. Thermodynamic parameters (Gibbs energy, enthalpy and entropy changes) were determined from the temperature dependence of K A for ([Cr(en)3]Cl3. The results of the study have been interpreted in terms of ion–solvent interactions and solvent properties.  相似文献   

15.
Quantum chemical calculations of structure and energies of various H-bonded complexes of phosphoric, phosphorous and methylphosphonic acids and their dimers with dimethylsulfoxide (DMSO), i.e., (acid) n –DMSO and acid–(DMSO) m for n = 1, 2 and m = 2, 3 have been carried out. The polar solvent effect is taken into account by using the CPCM model. It has been found that in DMSO environment the H-bonds in all complexes of investigated acid with DMSO are sizably stronger than the ones in the gas phase. At B3LYP-CPCM computation, the H-bonds between all investigated acid dimers and DMSO are significantly shorter than those found for complexes of corresponding acids with other compositions. The H-bonding interaction in acid–(DMSO) m for m = 1–3 becomes slightly weaker with increasing number DMSO molecules. The strength of the H-bond in all investigated complexes increases in the series of acids: (HO)2MePO < (HO)2P(O)H < H3PO4. Additionally, quantum theory of ‘atoms in molecules’ and natural bond orbitals method have been applied to analyze H-bond interactions.  相似文献   

16.
《Chemical physics》1987,116(1):101-111
The 2D and 19F spin-lattice relaxation times of 1,3,5-trifluorobenzene-d3 in 0.15 mol fraction solutions in various solvents have been measured over the temperature range 270–400 K. The relationship between the reorientational correlation times, τθ, and the angular momentum correlation times, τJ, obtained from the nuclear relaxation data has been compared with the theoretical relationships predicted by the J-diffusion limit of the extended diffusion (EDJ) model and the Fokker-Planck-Langevin (FPL) model for symmetric top molecules. It was found that the rotational motion of 1,3,5-trifluorobenzene-d3 in most solutions was better described by the FPL model with frictional anisotropy τ6· in the range 1–2. (τ6 and τ· are the correlation times for the angular velocity components along the axes parallel and perpendicular to the molecular symmetry axis.) The viscosity and temperature dependence of τθ has been analyzed in terms of a modified Debye equation and values for the anisotropic interaction parameter, κ, in each solvent are reported. The variation in κ with solvent is attributed mainly to the variation in the dipole moment of the solvent molecules. The frictional anisotropy (τ6·) is found to decrease as κ increases. This is interpreted in terms of the effects of molecular shape and electrostatic interactions between CF bond dipole moment and solvent dipole moments.  相似文献   

17.
Hydrochloric acid catalyzed hydrolysis of benzohydroxamic acid (C6H5CONHOH) has been studied in dioxane, acetone, DMSO, DMF, methanol, ethanol and isopropanol of varying compositions (10–70 vol.%) at 55°C. The anomalous behavior has been explained on the basis of solvent properties, solvation of the transition state and dielectric constant.  相似文献   

18.
The molecular structure of 1,1,1,2-tetrabromo-2,2-dimethyldisilane (Br3SiSiBrMe2) has been determined in the gas phase by electron diffraction and ab initio molecular-orbital calculations. The computational investigation was used to augment the experimental investigation using the Structure Analysis Restrained by Ab initio Calculations for Electron diffractioN (SARACEN) method. The structure was found to adopt a staggered structure with C s symmetry by both theory and experiment. Important structural parameters (r h1) include: rSi–Si 235.6(5) pm, rSi–C 185.4(3) pm, rSi–Brav 220.3(1) pm, ∠Si–Si–Br(14) 106.1(4)°, ∠Si–Si–C 109.2(8)° and ?Br–Si–Si–Br 180.0°(fixed). These experimental observations are supported by theoretical predictions obtained at the MP2/6-311+G* level. An analogous theoretical investigation was also performed for the series X3SiSiXMe2 (X = F, Cl and I) and structural trends identified. The Si–X bond was observed to lengthen as a function of the halogen substituent, with corresponding changes to the Si–Si–X bond angles in the SiX3 groups. The Si–Si–X bond angle in the SiXMe2 groups displayed rather different behaviour, and was relatively stable to substitution until X = I. The flexible nature of bond angles about silicon atoms was observed, even in this relatively sterically unhindered system.  相似文献   

19.
The effect of solvent on the strength of noncovalent interactions and ionic mobility of the dibenzo‐18‐crown‐6 complex with K+ in water/organic solvents was investigated by using affinity capillary electrophoresis. The proportion of organic solvent (methanol, ethanol, propan‐2‐ol, and acetonitrile) in the mixtures ranged from 0 to 100 vol.%. The stability constant, KKL, and actual ionic mobility of the dibenzo‐18‐crown‐6‐K+ complex were determined by the nonlinear regression analysis of the dependence of the effective electrophoretic mobility of dibenzo‐18‐crown‐6 on the concentration of K+ (added as KCl) in the background electrolyte (25 mM lithium acetate, pH 5.5, in the above mixed hydro–organic solvents). Competitive interaction of the dibenzo‐18‐crown‐6 with Li+ was observed and quantified in mixtures containing more than 60 vol.% of the organic solvent. However, the stability constant of the dibenzo‐18‐crown‐6‐Li+ complex was in all cases lower than 0.5 % of KKL. The log KKL increased approximately linearly in the range 1.62–4.98 with the increasing molar fraction of organic solvent in the above mixed solvents and with similar slopes for all four organic solvents used in this study. The ionic mobilities of the dibenzo‐18‐crown‐6‐K+ complex were in the range (6.1–43.4) × 10?9 m2 V?1 s?1.  相似文献   

20.
An indirect method is employed for determining the 15N parameters at the natural abundance level in a series of simple acyclic and cyclic amides. The one bond coupling constant, 1J(15N1H), and the 15N chemical shift are measured as a function of the carbonyl substituent group or the ring size and the nature of the solvent (CCl4 or H2O). These 15N parameters are related to the amide bond structure, the nitrogen configuration and possible intermolecular hydrogen bonding (amide-amide or amide-water).  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