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.
The first conformational analysis of 3‐silathiane and its C‐substituted derivatives, namely, 3,3‐dimethyl‐3‐silathiane 1 , 2,3,3‐trimethyl‐3‐silathiane 2 , and 2‐trimethylsilyl‐3,3‐dimethyl‐3‐silathiane 3 was performed by using dynamic NMR spectroscopy and B3LYP/6‐311G(d,p) quantum chemical calculations. From coalescence temperatures, ring inversion barriers ΔG for 1 and 2 were estimated to be 6.3 and 6.8 kcal/mol, respectively. These values are considerably lower than that of thiacyclohexane (9.4 kcal/mol) but slightly higher than the one of 1,1‐dimethylsilacyclohexane (5.5 kcal/mol). The conformational free energy for the methyl group in 2 (?ΔG° = 0.35 kcal/mol) derived from low‐temperature 13C NMR data is fairly consistent with the calculated value. For compound 2 , theoretical calculations give ΔE value close to zero for the equilibrium between the 2 ‐Meax and 2 ‐Meeq conformers. The calculated equatorial preference of the trimethylsilyl group in 3 is much more pronounced (?ΔG° = 1.8 kcal/mol) and the predominance of the 3 ‐SiMe3 eq conformer at room temperature was confirmed by the simulated 1H NMR and 2D NOESY spectra. The effect of the 2‐substituent on the structural parameters of 2 and 3 is discussed. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

2.
The conformational flexibility of three covalently linked dimers consisting of two xanthene‐based moieties connected by a diphenyl ether linker was studied using NMR spectroscopy, X‐ray crystallography, and density functional theory (DFT) calculations. The three dimers interconvert as a function of pH: the doubly cationic dimer (Xan+)2 exists in acidic solutions (pH < 0.5), the mono‐alcohol monocation Xan+–Xan‐OH at intermediate pH values (pH = 1–3), and the neutral diol at the highest pH‐values (pH > 3). Each dimer exhibits conformational degrees of freedom associated with rotations of either the xanthene moiety or of the diphenyl ether (DPE) linker. The barriers for rotation of the xanthylium moiety were evaluated using DFT calculations, yielding values of 23 kcal/mol for (Xan+)2 and 11 kcal/mol for (Xan‐OH)2, respectively. The rotational barrier for the diphenyl ether linker in Xan+–Xan‐OH (15 kcal/mol) was experimentally determined using variable temperature NMR measurements. The relative orientation of the two –OH groups in (Xan‐OH)2 diol was investigated in solution and the solid state using NMR spectroscopy and X‐ray crystallography. The conformer observed in the solid state was found to be the In–Out conformer, while free rotation of the xanthenol units is thought to occur on the NMR timescale at room temperature. These studies are relevant for the design of linkers for efficient water oxidation catalysts. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

3.
Extending our earlier findings for [3.3]paracyclophane, NMR line shape studies of the conformational dynamics in [3.2] and [4.3]paracyclophanes are reported, of which the former is conformationally homogeneous and the latter occurs in two enantiomeric forms. For [3.2]paracyclophane, the Arrhenius activation energy Ea = 11.6 ± 0.1 kcal/mol and preexponential factor log (A/s?1) = 12.92 ± 0.07 were found. In [4.3]paracyclophane, the conformational dynamics are quite complicated because, apart from interconversions of each enantiomer into itself proceeding via inversion of the propano bridge with rate constant k1, the enantiomers mutually rearrange with rate constant k2 due to inversion of the butano bridge. The determination of Arrhenius parameters from dynamic 1H spectra of the aromatic protons for these two conformational processes (Ea = 11.2 ± 0.5 kcal/mol and log (A/s?1) = 13.6 ± 0.5 for the former, and Ea = 9.7 ± 0.4 kcal/mol and log (A/s?1) = 13.2 ± 0.4 for the latter) is the highlight of this work. In the investigated temperature range, in [4.3]paracyclophane, the occurrence of other conformational processes beyond those mentioned above can be excluded, because they would produce different line shape patterns than those actually observed. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

4.
We here report on the conformational evolution of the bis(trifluoromethanesulfonyl)imide anion (TFSI) in protic and aprotic TFSI‐based ionic liquids as a function of temperature. The investigation is performed by Raman spectroscopy in the spectral ranges 240‐380 cm−1 and 715‐765 cm−1, where the interference from bands due to the cations is negligible. The contribution from each TFSI conformation, i.e. the cisoid (C1) and the transoid (C2), is quantified in order to estimate the enthalpy of conformational change, ΔH, which is found to be in the range 3.4–7.3 kJ/mol in the liquid state. Conformational information is for the first time determined from the 740 cm−1 band, which previously mainly has been used as an indicator of ion‐ion interactions. The similarity in ΔH values obtained from the two spectral ranges demonstrates the validity of using also the 740 cm−1 band for the quantification of the TFSI conformational evolution. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
The protonic motion in NH+4-H+(H2O)nβ″ -alumina wasstudied by measurements of proton spin-lattice and spin-spin relaxation times. The results show that two types of NH+4 reorientation, that is, three-fold reorientation about its unique C3 axis and tumbling motion as a whole, occur below room temperature. Above room temperature a fast proton translational diffusion was apparent, and its activation energy was estimated to be 36 kJ/mol. On the other hand, the conductivity measured on a pressed sample by an ac method is strongly humidity-sensitiveand changes by more than five orders from 0 to 100% relative humidity. The activation energy also may be humidity-sensitive. Its value was 73 kJ/mol for the sample equilibrated at 90°C in a dry nitrogen atmosphere.  相似文献   

6.
4,4‐Dimethyl‐1‐(trifluoromethylsulfonyl)‐1,4‐azasilinane 1 and 2,2,6,6‐tetramethyl‐4‐(trifluoromethylsulfonyl)‐1,4,2,6‐oxazadisilinane 2 were studied by variable temperature dynamic 1H, 13C, 19F NMR spectroscopy and theoretical calculations at the DFT (density functional theory) and MP2 (Møller‐Plesset 2) levels of theory. Both kinetic (barriers to ring inversion) and thermodynamic data (frozen conformational equilibria) could be obtained for the two compounds. The computations revealed two minima on the potential energy surface for molecules 1 and 2 corresponding to the rotamers with the CF3SO2 group directed ‘inward’ and ‘outward’ the ring, the latter being 0.2–0.4 kcal/mol (for 1 ) and 1.1 kcal/mol (for 2 ) more stable than the former. The vibrational calculations at the DFT and MP2 levels of theory give the values of the free energy difference ΔGo for the ‘inward’ ‘outward’ equilibrium consistent with those determined from the experimentally measured ratio of the rotamers. The structure of crystalline compound 2 was ascertained by X‐ray diffraction analysis. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

7.
The Raman and infrared spectra (4000 to 50 cm–1) of the gas, liquid or solution, and solid have been recorded of n‐propylamine, CH3CH2CH2NH2. Variable temperature (−60 to −100 °C) studies of the Raman (1175 to 625 cm–1) and far infrared (600 to 10 cm–1) spectra dissolved in liquid xenon were carried out. From these data, the five possible conformers were identified and their relative stabilities obtained with enthalpy difference relative to trans–trans (Tt) for trans–gauche (Tg) of 79 ± 9 cm–1 (0.9 ± 0.1 kJ/mol); for Gg of 91 ± 26 cm–1 (1.08 ± 0.3 kJ/mol); for Gg′ of 135 ± 21 cm–1 (1.61 ± 0.2 kJ/mol); for Gt of 143 ± 11 cm–1 (1.71 ± 0.1 kJ/mol). The percentage of the five conformers is estimated to be 18% for the Tt, 24 ± 1% for Tg, 23 ± 3% for Gg, 18 ± 1% for Gg′ and 18 ± 1% for Gt at ambient temperature. The conformational stabilities have been predicted from ab initio calculations utilizing several different basis sets up to aug‐cc‐pVTZ from both second‐order Møller–Plesset (MP2, full) and density functional theory calculations by the Becke, three‐parameter, Lee–Yang–Parr method. Vibrational assignments were provided for the observed bands for all five conformers, which are supported by MP2(full)/6‐31G(d) ab initio calculations to predict harmonic force constants, wavenumbers, infrared intensities, Raman activities and depolarization ratios for both conformers. Estimated r0 structural parameters were obtained from adjusted MP2(full)/6‐311+G(d,p) calculations. The results are discussed and compared with the corresponding properties of some related molecules. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

8.
Measurements of X-ray diffraction, electrical resistivity, and magnetization are reported across the Jahn–Teller phase transition in LaMnO3. Using a thermodynamic equation, we obtained the pressure derivative of the critical temperature (T JT ), dT JT /dP?=??28.3?K?GPa?1. This approach also reveals that 5.7(3)J?(mol?K)?1 comes from the volume change and 0.8(2)?J?(mol?K)?1 from the magnetic exchange interaction change across the phase transition. Around T JT , a robust increase in the electrical conductivity takes place and the electronic entropy change, which is assumed to be negligible for the majority of electronic systems, was found to be 1.8(3)?J?(mol?K)?1.  相似文献   

9.
Cold‐ and heat‐induced β‐lactoglobulin (BLG) transformations have been analyzed in the presence of 4 M urea, from Raman spectroscopy investigations carried out simultaneously in the low wavenumber range (10–400 cm−1) and in the amide I region (1500–1800 cm−1). These investigations show common features between the denaturation processes at low and high temperatures. The denatured states are reached via an intermediate state characterized by a soft tertiary structure without detectable conformational changes. This intermediate is intimately connected with a tetrahedral hydrogen‐bond structure of water which extends over a limited range. It is shown that the disruption of the hydrogen‐bond network of D2O has an important consequence on the solvent dynamics, which controls protein dynamics and is characterized by an anharmonic behavior. By monitoring the amide I mode, conformational changes are detected at low temperature (below 5 °C) and determined to be similar to those detected at high temperature in the presence of urea near 65 °C, and in the absence of urea near 80 °C. The conformational changes are described as a loss of α‐helix structures and a concomitant formation of β‐sheets. The temperature dependence of the amide I wavenumber in BLG dissolved in the 4 M urea aqueous solution was interpreted on the basis of a two‐state model, leading to the protein stability curve related to its molecular conformation. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

10.
We present a Raman spectroscopy investigation of the vibrational properties of l-histidine crystals at low temperatures. The temperature dependence of the spectra show discontinuities at 165 K, which we identify with modifications in the bonds associated to both the NH3 + and CO2 ? motifs indicative of a conformational phase transition that changes the intermolecular bonds. Additional evidence of such a phase transition is provided by differential scanning calorimetry measurements, which identified an enthalpic anomaly at ~165 K.  相似文献   

11.
The conformational analysis of the first representative of the Si‐alkoxy substituted six‐membered Si,N‐heterocycles, 1,3‐dimethyl‐3‐isopropoxy‐3‐silapiperidine, was performed by low‐temperature 1H and 13C NMR spectroscopy and DFT theoretical calculations. In contrast to the expectations from the conformational energies of methyl and alkoxy substituents, the Meaxi‐PrOeq conformer was found to predominate in the conformational equilibrium in the ratio Meaxi‐PrOeq : Meeqi‐PrOax of ca. 2 : 1 as from the 1H and 13C NMR study. The thermodynamic parameters obtained by the complete line shape analysis showed that the main contribution to the barrier to ring inversion originates from the entropy term of the free energy of activation. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

12.
The chemiluminescent emission from CH*, C*2, OH*, and CO*2 during the self-ignition of various mixtures of ethane with oxygen and argon behind reflected shock waves in the 1240–1790 K temperature range at a total concentration of the mixture М 5 = (1 ± 0.2) × 10?5 mol/cm3 is experimentally studied. It has been shown that the time-to-maximum in the emission intensity profiles is almost identical for all the emitters studied. How the pattern of the OH* emission profile changes with the temperature and mixture composition is examined. The CH* and C*2 emission profiles demonstrate virtually symbatic behavior in the covered ranges of temperature, pressure, and equivalence ratio. It is established that the emission signals from OH* and CO*2 appear earlier than the C*2 and CH* emission signals. The numerical simulation predictions are found to be in close agreement with the experimental results.  相似文献   

13.
A series of trans‐2‐aminocyclohexanol derivatives have been explored as powerful conformational pH triggers. On protonation of the amino group, a conformer with equatorial position of ammonio and hydroxy groups becomes predominant because of an intramolecular hydrogen bond and electrostatic interactions. The energy of these interactions was estimated to be above 10 kJ/mol and in some models exceeded 20 kJ/mol (strong enough to twist a ring in tert‐butyl derivatives). As a result of this conformational flip, all other substituents are forced to change their orientation. If the substituents are designed to perform certain geometry‐dependent functions, for example, as cation chelators or as lipid tails, such acid‐induced transition may be used to control the corresponding molecular properties. The pH sensitivity of conformational equilibria was explored by 1H nuclear magnetic resonance spectroscopy (NMR), and the titration curves were used for estimation of the pKa values of protonated compounds that varied from 2.6 to 8.5 (in d4‐methanol) depending on the structure of amino group. Thus, trans‐2‐aminocyclohexanols can be also used as conformational pH indicators in organic solvents.  相似文献   

14.
A comprehensive study of the relationship between the electronic specific heat coefficient () and the temperature square coefficient (A) of the electrical resistivity for a single, cubic, heavy fermion alloy system, UPt5-xAux is presented. In this alloy system, whose low temperature properties are consistent with the Fermi-liquid behavior, varies by more than a factor of 10 while the corresponding A coefficient changes by a factor larger than 200. A tracks changes in fairly well, but , postulated to have a universal value for heavy fermions, is not constant and varies from about 10-6 (x = 0, 0.5) to 10-5 cm (mol K/mJ)2 (x > 1.1), thus from a value typical of transition metals to that characteristic of other heavy fermion compounds. We have found a correlation between and magnetic characteristics such as the paramagnetic Curie-Weiss temperature and the low temperature magnetic susceptibility divided by . Received 29 January 1999  相似文献   

15.
The infrared spectra and stability of CO and H2O sorption over Ag-exchanged ZSM-5 zeolite were investigated by using density function theory (DFT). The changes of NBO charge show that the electron transfers from CO molecule to the Ag+ cation to form an σ-bond, and it accompanies by the back donation of d-electrons from Ag+ cation to the CO (π*) orbital as one and two CO molecules are adsorbed on Ag-ZSM-5. The free energy changes ΔG, −5.55 kcal/mol and 6.52 kcal/mol for one and two CO molecules, illustrate that the Ag+(CO)2 complex is unstable at the room temperature. The vibration frequency of C-O stretching of one CO molecule bonded to Ag+ ion at 2211 cm−1 is in good agreement with the experimental results. The calculated C-O symmetric and antisymmetric stretching frequencies in the Ag+(CO)2 complex shift to 2231 cm−1 and 2205 cm−1 when the second CO molecule is adsorbed. The calculated C-O stretching frequency in CO-Ag-ZSM-5-H2O complex shifts to 2199 cm−1, the symmetric and antisymmetric O-H stretching frequencies are 3390 cm−1 and 3869 cm−1, respectively. The Gibbs free energy change (ΔGH2O) is −6.58 kcal/mol as a H2O molecule is adsorbed on CO-Ag-ZSM-5 complex at 298 K. The results show that CO-Ag-ZSM-5-H2O complex is more stable at room temperature.  相似文献   

16.
流体静压力下快离子导体Rb4Cu16I7Cl13离子电导的研究   总被引:1,自引:0,他引:1       下载免费PDF全文
在流体静压力为0.5—11.6kbar,温度-80—80℃范围内,测量了Rb4Cu16I7Cl13多晶粉末“松散”样品和“致密”样品的离子电导。“松散”样品电导与压力的关系表明,在4.0—5.0kbar附近,电导存在极大值;“致密”样品电导随压力单调下降。在一定压力下,“致密”样品电导率随温度变化的趋势与常压结果相同,压力对Rb4Cu16I7Cl关键词:  相似文献   

17.
The ground-state bleaching of highly concentrated rhodamine 6G solutions in methanol is studied with intense picosecond light pulses. The ground-state recovery time changes with concentration from about 3.9 ns at low concentration (10–5 mol/dm3) to about 1 ps at high concentration (0.6 mol/dm3). The shortening of the absorption recovery time is determined by the concentration dependent quenching of theS 1-state population due to unbound dimers and by the intensity dependent longitudinal and transverse amplified spontaneous emission.  相似文献   

18.
L. Dai  H. Li  C. Liu  G. Su  S. Shan 《高压研究》2013,33(3):193-202
Electrical conductivities of pyroxenite were measured between frequencies of 10?1 and 106 Hz in a multi-anvil pressure apparatus using different solid buffers (Ni+NiO, Fe+Fe3O4, Fe+FeO and Mo+MoO2) to stabilize the partial pressure of oxygen. The temperature ranged from 1073 to 1423 K (800 to 1200 °C) and the pressure from 1.0 to 4.0 GPa. We observe that: (1) the electrical conductivity (σ) of pyroxenite depends on frequency; (2) σ tends to increase with rising temperature (T), and Log σ and 1/T obey a linear Arrhenius relationship; (3) under control of the buffer Fe+Fe3O4, σ tends to decrease with rising pressure, nevertheless the activation enthalpy tends to increase. For the first time we have obtained values for the activation energy and activation bulk volume of the main charge carriers, which are (1.60±0.07) eV and (0.05±0.03) cm3/mol, respectively; (4) for a given pressure and temperature, σ tends to rise with increased oxygen fugacity, whereas the activation enthalpy and preexponential factor tend to decrease; and (5) the behaviour of the electrical conductivity at high temperature and high pressure can be reasonably interpreted by assuming that small polarons provide the dominant conduction mechanism in the pyroxenite samples.  相似文献   

19.
The kinetics of the thermal polymerization of perfluoromethylvinyl ether (PFMVE) is studied at pressures of 3–13 kbar (300–1300 MPa) and temperatures of 80–260°C. The activation energy (E act = (76 ± 3) kJ/mol) and activation volume (ΔV0 = −(27 ± 2) cm3/mol) for the overall polymerization rate are determined. The inhibition method is used to estimate the activation energy of thermal initiation (E in = (79.9 ± 3) kJ/mol). The quantity E p − (1/2)E t was calculated to be 36.6 ± 3 kJ/mol. The limiting polymerization temperature was evaluated: T lim = (180 ± 3)°C. A mechanism of PFMVE polymerization is proposed on the assumption that the reaction is bimolecular.  相似文献   

20.
Photoluminescence and excitation spectra of the spinel-type MgGa2O4 with 0.5 mol. % Mn2+ ions and Eu3+ content from 0 to 8 mol. % have been investigated in this work at room temperature. Polycrystalline samples were synthesized via high-temperature solid-state reaction method. Photoluminescence spectra of all samples exhibit host emission presented by a broad “blue” band peaking ∼430 nm, which consists of at least three elementary bands that correspond to different host defects. Excitation of the host luminescence showed the broad band with a maximum at 360 nm. Characteristic bands of d–d transitions of Mn2+ ions and f–f transitions of Eu3+ ions together with charge-transfer bands (CTB) of these ions were also found on the excitation spectra. Mn2+ and Eu3+ co-doped samples emit in green and red spectral regions. Mn2+ ions are responsible for the green emission band at 505 nm (4Т16А1 transition). The studies of photoluminescence spectra of activated samples with different Eu3+ ions content show characteristic f–f luminesecence of Eu3+ ions. The maximum of Eu3+ emission was found at 618 nm (5D07F2) and optimal concentration of activator ions was around 4 mol. %.  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