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.
Infrared spectroscopy studies of 2-acetylthiophene (ACTH) in 18 different organic solvents, both polar and non-polar, were undertaken to investigate the solvent-solute interactions. The frequencies of carbonyl stretching vibration upsilon(C=O) of ACTH were correlated with the properties such as the solvent acceptor number (AN) and the linear solvation energy relationships (LSER). The solvent-induced stretching vibration frequency shifts showed a better correlation with the LSER than the AN. A six-membered ring-like hydrogen bonding structure was presented and the solvent effects of ACTH in alcohol solvents were investigated in detail.  相似文献   

2.
The infrared spectroscopy studies of the C3 and C20 carbonyl stretching vibrations (upsilon(C=O)) of progesterone in CHCl3/cyclo-C6H12 binary solvent systems were undertaken to investigate the solute-solvent interactions. With the mole fraction of CHC13 in the binary solvent mixtures increase, three types of C3 and C20 carbonyl stretching vibration band of progesterone are observed, respectively. The assignments of upsilon(C=O) of progesterone are discussed in detail. In the CHCl3-rich binary solvent systems or pure CHCl3 solvent, two kinds of solute-solvent hydrogen bonding interactions coexist for C20 C=O. Comparisons are drawn for the solvent sensitivities of upsilon(C=O) for acetophenone and 5alpha-androstan-3,17-dione, respectively.  相似文献   

3.
The surface tension of binary solvents is modelled by analogy to solvation effects arising from solvent-solute interactions. Competitive exchange equilibria are postulated between solvent component I (water) and solvent component 2 (organic cosolvent) for solute, which in this case is air; the solvation shell is thus the surface phase. A quantitative relationship is given between surface tension and mole fractions x1 and x2, the model parameters being exchange equilibrium constants K1 and K2. The equation is analyzed, it is applied to literature surface tension data, and it is compared with an earlier model from this laboratory. Curve-fits are very good, and the parameters appear to possess physical significance.  相似文献   

4.
Research of methyl methacrylate (MMA) in three kinds of binary solvent systems (CCl4/C6H14, CHCl3/C6H14 and C2H5OH/C6H14) on the infrared (IR) spectra was reported. Two types of carbonyl stretching vibration bands for MMA in CHCl3/C6H14 or C2H5OH/C6H14 mixtures were found with the changing of the mole fraction of CHCl3 (XCHCl3) or C2H5OH (XC2H5OH). The carbonyl stretching vibration bands at lower frequencies in the above two mixtures were attributed to the formation of hydrogen bonding between MMA and CHCl3 or C2H5OH. While in CCl4/C6H14 mixtures there was only one type of carbonyl stretching vibration band of MMA. Good linear correlations between the frequencies of C=O or C=C stretching vibration band of MMA and XCCl4, XCHCl3 or XC2H5OH were found, respectively. The solute-solvent interactions in the three different binary solvent systems were discussed in detail.  相似文献   

5.
Infrared spectroscopy studies of six beta-alkoxyvinyl methyl ketones, with common structure R(1)O-CR(2)CH-COR(3), where R(1)=R(3)=CH(3), R(2)=H (1); R(1)=C(2)H(5), R(2)=H (2); R(3)=CF(3); R(1)=R(2)=CH(3), R(3)=CF(3) (3); R(1)=C(2)H(5), R(2)=C(6)H(5), R(3)=CF(3) (4); R(1)=C(2)H(5), R(2)=4-O(2)NC(6)H(4), R(3)=CF(3) (5); R(1)=C(2)H(5), R(2)=C(CH(3))(3), R(3)=CF(3) (6) in 11 pure organic solvents of different polarity were undertaken to investigate the solute-solvent interactions and to correlate solvent properties by means of linear solvation energy relationships (LSER) with the carbonyl and vinyl stretching vibrations of existing stereoisomeric forms. It was shown that contrary to simple carbonyl-containing compounds where solvent HBD acidity (alpha) has the largest influence on the nu (CO) band shift to lower wavenumbers, the dipolarity/polarizability (pi) term plays the main role in the interactions of conjugated enones with solvent molecules leading to the nu (CO) and nu (CC) bathochromic band shifts. The trifluoroacetyl group possesses a reduced ability to form hydrogen bonds with solvents. For the nu (CC) band of non-fluorinated enone 1 solvent HBD acidity (alpha) and solvent HBA basicity term (beta) play a perceptible role, whereas for 2 these terms are not significant. beta-Substituents in fluorinated enones such as R(2)=H, C(6)H(5), and C(CH(3))(3) assist in the intermolecular hydrogen bond formation of the carbonyl moiety with HBD solvents, while beta-substituents such as CH(3) and 4-NO(2)C(6)H(4) prevent the CO group to form the H-bonds with HBD solvents (the solvent HBD acidity term (alpha) is not significant). The comparison of four conformers of the enone 1 reveals that (EEE) form is the most polarizable conformer; the influences of the solvent dipolarity/polarizability (pi) and solvent HBD acidity (alpha) term on the bathochromic nu (CO) band shift are opposite to one another.  相似文献   

6.
The thermal decomposition reaction of acetone cyclic triperoxide, acetone cyclic diperoxide, 4‐heptanone cyclic diperoxide, and pinacolone cyclic diperoxide ca. 0.02 M was studied in pure solvents (acetone and 1‐propanol) and in binary mixtures of acetone/1‐propanol at 150°C. The kinetics of each system was explored by gas chromatography (GC) at different solvent compositions. The reactions showed a behavior accordingly with a pseudo‐first‐order kinetic law up to at least 90% peroxide decomposition. The main organic products derived from these thermolysis reactions were detected by GC analysis. Among them, the corresponding ketones, methane, ethane, and propane were the main identified products. The rates of decomposition of pinacolone diperoxide in the pure solvents were practically independent of the solvent characteristics, so it was of no interest to analyze its kinetic behavior in binary solvent mixtures. In acetone/1‐propanol mixtures, the solvation effect on the cyclic peroxides derived from 4‐heptanone and acetone molecules was slightly dominated by specific interactions between 1‐propanol and a diradical‐activated complex initially formed. This species was preferentially solvated by 1‐propanol instead of acetone. Specific interactions between the O atoms from the peroxidic bond and the H from the OH in 1‐propanol can be taken into account. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 347–353, 2010  相似文献   

7.
Infrared spectra of thionyl and seleninyl chlorides are reported and discussed. Complete assignment for both molecules is given on the basis of Cs symmetry. Solvent effects on XO bond (X = S, Se) have been studied in comparison with CO bond for carbonyl compounds and by means of the Kirkwood, Bauer and Magat relationship. Acceptor properties of oxychlorides for pyridinic adducts are shown.  相似文献   

8.
The standard free energy change for complex formation is written as a sum of effects arising from solvent-solvent interactions (the general medium effect), solvent-solute interactions (the solvation effect), and solute-solute interactions (the intersolute effect). The general medium effect is given by gA(–o), where g is a curvature correction factor to the solvent surface tension , A is the change in surface area as the two solvent cavities containing the substrate (naphthalene) and ligand (theophylline) collapse into a single cavity containing the complex, and o is the value of surface tcnsion at which there is no net solvophobic interaction; is defined to be the value appropriate to the equilibrium mean solvation shell composition. The solvation effect is modeled by equilibrium stoichiometric formation of solvated species. All data are related to the fully aqueous system to give MGo, the solvent effect on the free energy change, as an explicit function of solvent composition. Solvent effects on bimolecular association are related to solvent effects on the solubilities of the substrate, ligand, and complex. Approximation methods for interpreting such systems are described and are applied to the naphthalene-theophylline complex. It is shown that complex destabilization in mixed aqueous-organic solvents (relative to the fully aqueous system) may receive contributions from both the general medium and the solvation effects, and that these contributions can be quantitatively estimated.  相似文献   

9.
10.
11.
The fluorescence steady-state emission spectra of lipophilic fluorescence probe PRODAN in ethanol/buffer solvents of different concentrations (0.3, 0.9, 3 mol L(-1) ethanol) were extensively studied and analytically described. The complex experimental spectra, corrected for background effects, were fitted by two Gaussian curves. The energy separation of two maxima, (0.147+/-0.002) eV at 37 degrees C and (0.143+/-0.003) eV at 25 degrees C, was independent of ethanol concentration. The blue shifts observed for both maxima were linearly dependent on solvent polarity. The linear dependences of fluorescence's intensities on PRODAN concentration in all ethanol/buffer solvents indicate that no PRODAN self-quenching takes place even at the highest measured PRODAN concentrations.  相似文献   

12.
The charge-transfer transition energies of 2,6-dimethoxynaphthalene and 9-methylanthracene with tetrachlorophthalic anhydride, and acenaphthene and anthracene with 3,5-dinitrophthalic anhydride were measured in sixty aprotic solvents. The observed effects can be interpreted in terms of various solvent parameters if the solvents are divided into the following classes: halogen-containing, aromatic and n-donor solvents.  相似文献   

13.
14.
NMR solvent effects induced by benzene in methyl substituted benzoic esters can be effectively used in the interpretation of complex NMR spectra. A 1:1 collision complex between solvent and solute is proposed, supported by dilution curve experiments and the geometry of the ‘complex’ is discussed.  相似文献   

15.
The NMR spectra of stereoblock poly(methyl methacrylate) in several solvents were measured. It is concluded from the following experimental results that the solute–solvent complexes are formed in benzene solution: the chemical shifts measured in C6H6 go to a lower field than do those in CDCl3, except those of the ester methyl group, which splits into three resonances, and the shifts in the aromatic solvents are so different from those in the aliphatic solvents that Buckingham's theory cannot be applied to the results. The analysis of the temperature dependence of the chemical shifts of PMMA in benzene solution gave the heat of formation of the complex: ΔH = 2.8 ± 0.5 kcal./mole.  相似文献   

16.
《Chemical physics letters》1986,127(4):387-391
The pure radiative lifetime of ovalene was measured in a series of solvents. The effect of S1–S2 energy gap reduction by the solvent on the pure radiative lifetime was studied and could be fitted to the theoretical calculations. Several questions are raised concerning the validity of the theory at small inter-electronic energy gaps.  相似文献   

17.
The NMR spectra of pyridine, pyrazine and imidazole are examined in acetone-d6. The differential shifts of the ring protons are accounted for in terms of an interaction between the lone pair on the ring nitrogen atom and the carbonyl group of acetone.  相似文献   

18.
Summary The chemiluminescent reaction of lucigenin with basic hydrogen peroxide has been studied in several mixtures of water with the cosolvents methanol, ethanol, 1-propanol, dimethylformamide, and dimethylsulfoxide. The chemiluminescence spectra depend on the cosolvent and its concentration in the reaction medium. With increasing cosolvent concentration, the chemiluminescence shifts to lower wavelengths. For similar cosolvents, the size of this shift increases with decreasing dielectric constant. In high-cosolvent-concentration mixtures, the chemiluminescence matches the fluorescent emission of N-methylacridone. Chemiluminescence from low-cosolvent-concentration mixtures is explained as the sum of the lucigenin and N-methylacridone fluorescent emissions, the lucigenin emission probably being a consequence of energy transfer from N-methyl-acridone. The cosolvent inhibits this energy transfer. These observations, taken together with our previous kinetic results, indicate that the reaction mechanism is the same in all the studied reaction media.
Lösungsmitteleffekte bei der Reaktion von Lucigenin mit basischem Hydrogenperoxid: Chemilumineszenzspektren in gemischten polaren Lösungsmitteln
Zusammenfassung Die Chemilumineszenzreaktion von Lucigenin mit basischem Hydrogenperoxid wurde in verschiedenen Mischungen von Wasser mit Methanol, Ethanol, 1-Propanol, Dimethylformamid oder Dimethylsulfoxyd untersucht. Die Chemilumineszenzspektren hängen vom organischen Kosolvens und dessen Konzentration im Reaktionsmedium ab. Mit ansteigender Konzentration ergeben sich in der Chemilumineszenz Verschiebungen zu größeren Wellenlängen. Für ähnliche Kosolventien steigt diese Verschiebung mit kleineren Dielektrizitätskonstanten an. Bei hohen Kosolvenskonzentrationen gleicht die Chemilumineszenz der Fluoreszenzemission von N-Methylacridon. Die Chemilumineszenz bei kleinen Kosolvenskonzentrationen kann als die Summe der Fluoreszenzemission von Lucigenin und N-Methylacridon erklärt werden, wobei die Lucigeninemission vermutlich eine Folge eines Energietransfers von N-Methylacridon ist. Das Kosolvens verhindert diesen Energietransfer. Diese Beobachtungen, zusammen mit früheren kinetischen Resultaten, erlauben den Schluß, daß der Reaktionsmechanismus in allen Reaktionsmedien der gleiche ist.
  相似文献   

19.
NMR solvent effects induced by aromatic solvents on some 1,4-dioxanes, 1,3-dioxolanes and on some sulphur analogue derivatives are reported. The shielding effect of the aromatic solvents is examined in respect to the structure of the solute.  相似文献   

20.
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