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 46 毫秒
1.
目的:建立刺激胰岛素分泌的新型降糖药物(-)-2 (S)-苄基-4-酮-4-(顺式-全氢化异吲哚-2-基)丁酸钙对映体的HPLC拆分方法。方法:采用Sumichiral OA-3300手性柱(250 × 4.6 mm I.D., 5 μm), 柱温35℃,以0.05 mol·L-1醋酸铵的甲醇溶液为流动相,检测波长为210 nm。结果:本品两对映体在22分钟内实现良好分离,分离度达3以上,S-异构体分别在0.028 ~ 5.6 μg mL-1和0.03 ~ 6.0 μg mL-1范围内线性关系良好,回归方程分别为:Y=1.32×103x-2.54 (r=0.9997)和Y=1.15×103x-1.78 (r=0.9998),最低检测限分别为0.15 ng和0.10 ng,方法精密度RSD低于1.0% (n=5)。结论:建立的对映体分离方法可用于本品光学异构体的质量控制。  相似文献   

2.
Semiempirical (AM1 and PM3) and density functional theory (DFT) calculations were performed on about 50 porphyrin isomers with 25 each of 1,2 (syn) and 1,3 (anti) tautomeric forms. The corresponding oxa-and thia-core-modified analogues were also computed. The variations of relative energies and stabilities of the core-modified analogues were compared with parent porphyrin1 and the corresponding oxa-and thia-analogues. The trends in relative energies are not significantly changed while going from parent system to oxa-and thia-core-modified porphyrins in case of bothsyn andanti tautomers. Isomers of types [2·2·0·0], [3·0·1·0], [3·1·0·0], and [4·0·0·0] are destabilized due to the absence of methine bridge, which results in angle strain for tetrapyrroles. Isomers having [2·1·1·0], [2·1·0·1], [2·0·2·0] and [2·2·0·0] connectivity, the Z isomers, are more stable compared to the correspondingE isomers in bothsyn andanti forms of parent and core-modified analogues.  相似文献   

3.
The chiral complex, (4S,11S)-[Cu(1,7-CT)(μ-CN)Fe(CN)4NO] ·?H2O (1,7-CT =?5,5,7,12,12,14-hexamethyl-1,4,8,11-tetraazacyclotetradeca-1,7-diene), and its enantiomer have been synthesized by reaction and conglomerate crystallization. They consist of heterobinuclear species in which the Cu and Fe centers are linked by a cyanide bridge and crystalline water. The Cu(II) is coordinated by five N atoms and exhibits a distorted square-pyramidal geometry, in which two hydrogen atoms on secondary amines lie in the inward side of the macrocyclic plane, while on the other moiety the Fe(II) is a slightly distorted octahedral structure. The binuclear molecules are linked through intermolecular O2–H2A···N1 and O2–H2B···N4 hydrogen bonds, forming two different waved chains that oriented the molecules for optical activity. IR spectrum shows the existence of bridging cyanide ligand. In methanol the specific rotations of enantiomers are ±205 deg ·?cm2 ·?(10 g)?1, the peak positions of their circular dichroism spectra are close to that of their UV-Vis spectra and present up and down symmetric signals.  相似文献   

4.
A simple liquid chromatographic method was developed for the separation and quantification of voriconazole and its enantiomer in drug substance. The separation was achieved on Chiral cel-OD (250 mm × 4.6 mm × 10 μm) using mobile phase consisting of n-hexane and ethanol in the ratio 9:1 (v/v) with a flow rate of 1.0 mL min−1, at 27 °C column temperature and detection at 254 nm with an injection volume of 20 μL. Ethanol was used as diluent. The method is capable of detecting the (2S, 3R) enantiomer down to 0.0075% and can quantify down to 0.021% with respect to sample concentration. The method is rapid and the resolution achieved was about 3.0. This method can be employed for the quantification of (2S, 3R) enantiomer in voriconazole drug substance.  相似文献   

5.
 It was found that the hypericinate salts of (R)-1-phenylethylamine and (S)-1-(1-naphthyl)-ethylamine display a small chiroptical signal of the same sign only at high concentrations in an apolar solvent. No further indications of a chiral discrimination between the helical conformers of hypericinate could be found in these cases. However, upon esterification of the 3-hydroxyl group of hypericin with (1S)-camphanic chloride, the two diastereomers were found in an 1:1 ratio equilibrating rather fast at temperatures above 30°C with one diastereomer in excess. From the temperature dependence of the equilibrium positions (measured by means of CD and 1H NMR), a ΔG 0 value of 5.8±0.5 kJ·mol−1 was derived. Accordingly, the chiral discrimination of the (M)-configured enantiomer of the helix by the (S)-configured auxiliary occurred at an intermediate level. From the temperature dependence of the equilibration kinetics an activation energy of E a = 70±0.5 kJ·mol−1 was derived, which thus defines the upper limit of the helix inversion of hypericin and hypericinate. This value is by about 10 kJ·mol−1 lower than the recently estimated limit.  相似文献   

6.
The temperature dependence of the NMR spectrum of 5,5-dimethyl-3,7-dithia-1,2-benzocyclo-heptene-( 1 ) is described and discussed. This compound occurs in two conformers with different topographies of the seven membered ring. From the chemical shift of the 2 geminal methyl groups (obtained by low-temperature spectra) it can be shown that the ring occurs in one case in the chair form and in the other case in the twist form. The free conformational energy of the twist form is only about 20 cal/mole. Two conformational changes can be distinguished: the conversion between chair and twist forms and the pseudorotation of the twist forms. The free activation enthalpies of the conversion (ΔGV) and the pseudorotation (ΔGP) are 12·6 ± 0·1 Kcal/mole (at ?12°C) and 8·3 ± 0·3 Kcal/mole (at ?95°C) respectively.  相似文献   

7.
The ability of capillary zone electrophoresis in the development of analytical methods devoted to the quality control of the thiol drug penicillamine is shown. Using 50 mM phosphate running buffer (pH 2.5), good quantitations of underivatized penicillamine and its disulfide were achieved; detection at 200 nm allowed checking the presence of the disulfide impurity in pharmaceuticals. The use of 1,1-[ethenylidenebis(sulfonyl)]bis-benzene as a thiol specific reagent resulted in an increased sensitivity for the quantitation of D-penicillamine (limit of detection at 200 nm wavelength was 1.5 microM). Introducing beta-cyclodextrin as chiral selector in the running buffer, enantioseparation of D-L-penicillamine was obtained; for this purpose (+)-camphor-10-sulfonic acid, a chiral ion-pairing reagent, was found to be an essential additive in obtaining a baseline separation. The resulting enantioseparative system was validated in order to evaluate the presence of the toxic L-penicillamine enantiomer in pharmaceutical samples.  相似文献   

8.
The NMR spectra of 1,1-dichloro-2,2-difluoroethane ( 1 ), 1,1-dibromo-2,2-difluoroethane ( 2 ), meso and dl 1,2-dichloro-1,2-difluoroethane ( 3 ) and 1,1,2,2-tetrachloroethane ( 4 ) have been analysed in a number of solvents. The 19F spectrum of 3 in L-bornyl acetate at 56·4 and 94·1 MHz allows an unambiguous identification of the meso and dl isomers. The spectra of the d and l isomers consist of two AA′XX′ spectra with a small chemical shift difference between the d and l forms, whilst that of the meso form is an apparent AA′XX′ spectrum at 56·4 MHz but an ABXX′ spectrum at 94·1 MHz, the 19F nuclei in this isomer being anisochronous in this solvent. The observed solvent and temperature dependence of the couplings of 1, 2 , meso 3 and 4 when combined with the calculated solvation energies, allow the determination of the rotamer energies and couplings in these molecules. The rotamer energy differences (EgEt) in the liquid and vapour states are 0·6 and ?0·2 kcal/mol ( 1 ); 0·4 and ?0·5 kcal/mol ( 2 ); 0·9 and 0·2 kcal/mol meso ( 3 ) and ?0·1 and ?0·8 kcal/mol ( 4 ). The 3J(HH), 3J(HF) and 3J(FF) couplings for the distinct rotamers are considered together with those of similarly constituted molecules. The general agreement demonstrates that the solvation theory may be applied to multisubstituted ethanes without any basic modifications. The trans oriented HH couplings show a linear substituent electronegativity dependence, which differs appreciably from that obtained for disubstituted ethanes, however. The gauche couplings show the influence of dihedral angle variations as well as substituent electronegativity. The rotamer 3J(FF) couplings in meso 3 are ?38·2 Hz (Jt) and ?17·4 Hz (Jg).  相似文献   

9.
The suitability of n-hexane and n-octane as reactant gases in chemical ionization mass spectrometry has been investigated. The mass spectra of these substances have been investigated as a function of pressure up to 2·4 Torr for n-hexane and 1·7 Torr for n-octane. The major ion present in n-hexane at 0·8 Torr is [C6H13]+ (m/e 85) with a relative intensity of 0·65. In n-octane at 0·8 Torr the major ions are [C8H17]+ (m/e 113), [C6H13]+ (m/e 85) and [C5H11]+ (m/e 71). The relative intensities of these ions are 0·38, 0·12 and 0·19, respectively. These alkyl ions in both n-hexane and n-octane are thought to have tertiary structures. Rate constants for the rates of reaction of the primary ions in the two compounds have been determined. The n-hexane chemical ionization spectra of 26 compounds were determined. The spectra of polar compounds are dominated by proton transfer, whereas those of nonpolar compounds exhibit proton transfer and in addition often surprisingly large amounts of electron transfer. The n-octane chemical ionization spectra of 15 compounds were determined and the spectra in general are quite similar to those obtained with n-hexane. n-Hexane and n-octane can be used as reagents in analytical chemical ionization mass spectrometry, but except in certain specialized uses they would probably have no advantage over i-butane.  相似文献   

10.
Data for 30 hydrogen bonding pairs taken from the alkanethiols, i-C3H7SH, nC3H9SH and t-C4H9SH, and 16 bases have been obtained by a PMR method. Representative data for i-C3H7SH at 304 ± 2°K are (base, 102K in M?1, –ΔH° in kcal/mole): (CH2)4S, 3·1, 0·8; (CH3)2S, 3·0, 0·9; (CH3)2S2, 3·7, 0·5; (CH3)2CO, 4·7, 0·9; CH3COOC2H5, 5·7, 1·1; (CH2)4O, 6·1, 1·0; HCON(CH3)2, 12, 0·9; (CH3 O)2 SO, 12, 0·9; (C2 H5O)3PO, 6·5, 1·0; CH3 (CH3 O)2PO, 18, 1·0; ((CH3)2N)2 CO, 5·9, 1·1; CH3 CN, 13, 0·6. In essence, the problems and errors involved in obtaining equilibrium data for weak complexes stem from the limited concentration rangethat is accessible. This leads to large uncertainties in the quantities K, ΔH° and ΔS°. Structural effects on hydrogen bonding at the sulfur site, both as Lewis acid or base, are discussed. Two erroneous PMR methods in the literature used for assessing the strength of hydrogen bonds are pointed out.  相似文献   

11.
The photogalvanic effect has been studied in three systems using photogalvanic cells and NaLS-ascorbic acid-azur A, NaLS-ascorbic acid-azur B, NaLS-ascorbic acid-azur C systems. The photopotential and photocurrent generated by these systems are 7700, 9710, 623·0 mV and 160·0, 185·0, 145·0 (μA respectively. The effects of different parameters on the electrical outputs of the cell have been observed and current-voltage characteristics of the cell studied, and a mechanism has been proposed for the generation of photocurrent in photogalvanic cells. The conversion efficiencies for azur A, azur B and azur C are 0·5461, 0·9646 and 0·4567% and storage capacity 110, 135 and 95 min respectively.  相似文献   

12.
The carboxyl proton chemical shifts of neat trimethylacetic acid and its solutions in cyclohexane have been measured as a function of temperature. Formation, at the melting point, of a carboxyl group satellite line on the low field side has been found. The intensity of this satellite line rises with decreasing temperature, whereas the intensity of the main line decreases. The chemical shifts of the monomer (δm = 4·9 ± 0·6 ppm) and dimers, the equilibrium constants, enthalpy (ΔH = ?(11·3 ± 0·6) kcal/mol) and entropy (ΔS = ?(13·5 ± 0·7) cal/mol. degree) changes have been calculated using the concentration dependence of the chemical shifts of trimethylacetic acid in cyclohexane solutions at various temperatures. The chemical shift of the satellite line has been shown to correspond to the chemical shift of the cyclic dimers of the acid.  相似文献   

13.
Amide-rotational barriers in some 2-substituted N,N-dimethylcyclopropanecarboxamides in CDCl3 as a solvent (0·25 M) were obtained with an iterative total line shape analysis. N,N-dimethylcyclopropanecarboxamide shows a barrier ΔG = 16·72 ± 0·01 kcal/mole. Para-nitro substitution in (trans)-2-phenyl-N,N-dimethylcyclopropanecarboxamide raises the barrier ΔG from 17·08 ± 0·01 to 17·40 ± 0·02 kcal/mole.  相似文献   

14.
Abstract

The conformations of flexible chain molecules incorporated in a nematic environment have been investigated. The phase behaviours and orientational characteristics of 1,2-dimethoxyethane (DME) dissolved in 4′-methoxybenzylidene-4-n-butylaniline at low solute mol fraction have been reported in our previous paper. In this work, proton–proton and carbon–carbon dipolar coupling constant measurements were attempted in addition to 2H NMR observations of quadrupolar splittings. These conformation-dependent properties were analysed according to the rotational isometric state (RIS) simulation scheme previously proposed. Our treatment rests on the assumption that the molecular axis of the chain should tend to align along the nematic field. Studies were further extended to a mixture of 1,2-diphenyloxyethane (DPE) with a nematic liquid crystal, 4,4′-azoxyanisole. Replacement of the terminal methyl groups in DME by phenyl groups leads to DPE. The results of the analysis indicate that the fraction of elongated conformers such as ttt tends to increase significantly on going from the isotropic to the liquid crystalline solution: DME; 9·4 to 15·6 per cent and DPE; 12·3 to 20·0 per cent. The tgt form (tg + t or tg ? t) is the lowest-energy arrangement of these molecules. The fraction of this conformer also increases in DPE (26·8 to 40·0 per cent), while it decreases in DME (25·7 to 18·8 per cent). The conformational distributions of chain molecules were found to remain invariant over the range of concentration (0·5–6·0 mol per cent) and temperature (22·0–44·5°C) studied. It has been concluded that the observed variation of proton dipolar and deuterium quadrupolar couplings with concentration and temperature mainly arises from the orientational order of the molecular axis, which varies sensitively with the alignment of the surrounding solvent molecules. These results suggest that flexible chains are also participating in the nematic interaction by adjusting their configurations so as to enhance favourable interactions and suppress unfavourable steric repulsions when accommodated in an anisotropic potential field.  相似文献   

15.
The chemical shifts of the aldehydic proton in furfural, thiophenealdehyde and benzaldehyde have been measured in fourteen solvents. The correlation of the chemical shifts of thiophenealdehyde and benzaldehyde is excellent (r = 0·996) while it is lower for furfural–benzaldehyde (r = 0·956). The long range coupling constant Jα5 of furiurai has been measured in twelve solvents and the rotameric mole fractions determined. The chemical shifts of individual rotamers are calculated and shown to correlate with benzaldehyde, (r = 0·992; 0·993). Only one rotamer is predominant for thiophenaldehyde in all solvents. The intrinsic solvent effects of the three aldehydes are similar.  相似文献   

16.
Summary.  It was found that the hypericinate salts of (R)-1-phenylethylamine and (S)-1-(1-naphthyl)-ethylamine display a small chiroptical signal of the same sign only at high concentrations in an apolar solvent. No further indications of a chiral discrimination between the helical conformers of hypericinate could be found in these cases. However, upon esterification of the 3-hydroxyl group of hypericin with (1S)-camphanic chloride, the two diastereomers were found in an 1:1 ratio equilibrating rather fast at temperatures above 30°C with one diastereomer in excess. From the temperature dependence of the equilibrium positions (measured by means of CD and 1H NMR), a ΔG 0 value of 5.8±0.5 kJ·mol−1 was derived. Accordingly, the chiral discrimination of the (M)-configured enantiomer of the helix by the (S)-configured auxiliary occurred at an intermediate level. From the temperature dependence of the equilibration kinetics an activation energy of E a = 70±0.5 kJ·mol−1 was derived, which thus defines the upper limit of the helix inversion of hypericin and hypericinate. This value is by about 10 kJ·mol−1 lower than the recently estimated limit. Corresponding author. E-mail: heinz.falk@jku.at Received March 22, 2002; accepted April 3, 2002  相似文献   

17.
Abstract— The spectra of absorption, fluorescence and excitation of monolayers and thin films containing chlorophyll a together with a carotenoid (cis-β-carotene, trans-β-carotene, fucoxanthin, or zeaxanthin), were measured at — 196°C. The concentration ratios used, (Chl)/(Car), were 6:1, 4:1, 3:1, 2:1, 1:1 and 1:3, and the area densities, 3·70, 2·55, 1·76, 0·71, 0·37 and 0·17 nm2/pigment molecule. In dilute monolayers, (3·70 nm2/molecule), with a constant concentration ratio (Chl)/(Car) = 3:1, evidence of three β-carotene forms, with absorption bands at 460, 500 and 520 nm (C460, C500 and C520), and of a chlorophyll a form with an absorption band at 669–672 (Chl669–672) was found. On increasing the density to 0·2–0·3 nm2/molecule, a conversion of C460 and C520 into C500, was observed, and several more additional (probably more strongly aggregated) chlorophyll a forms appeared, with absorption bands at 672–733 nm. With excess carotene [(Chi)/(Car) = 1:3] the forms C460, C500, C520 and Chl669–672 were present even in the most dense films (0·2–0·3 nm2/molecule). The same was found with other carotenoids: if one of the pigments was in excess, aggregated forms of the other tended to disappear. In the transfer of energy from carotenoids to chlorophyll a, C500 was found to be the main donor. In layers with a concentration ratio (Chl)/(Car) = 3:1, the efficiency of transfer was less than 10 per cent at the lowest density used (3·70 nm2/molecule); it increased to 50 per cent, as the density was increased to 0·20 nm2/molecule. When the relative concentration of the carotenoid was increased to (Chl)/(Car) = 1:1, the efficiency of energy transfer dropped to 25 per cent even at 0·20 nm2/molecule. It seems that the efficiency of energy transfer between carotene molecules (prior to its transfer to chlorophyll a) is low, and effective transfer occurs only between β-carotene and immediately adjacent chlorophyll a molecules.  相似文献   

18.
    
Hirudonine sulphate (C9H23N7. 1·5 H2SO4. 2·5 H2O) is triclinic inPI space group with cell constantsa=7·168(9),b=14·534(6),c=11·918(5) ?, α=110·50(3), β=108·75(6) and γ=79·16(6)°,V=1097(2)?3,Mr=421·4,Z=2,d x=1·358(2) gcm−3,d c=1·276 gcm−3. MoKα (λ=0·7903 ?), μ=1·94 cm−1,F(000)=436,T=295 K,R(F)=0·144. The structure was solved by direct methods and refined to a final R factor of 0·144 for 1036 unique reflections. One of the sulphur atoms is in special position and is disordered. The amine molecule is hydrogen-bonded to the sulphate oxygen through water molecules. Water channels are formed at unique places involving water oxygens, amine and sulphate oxygens along thea axis. DCB contribution Number 712.  相似文献   

19.
Abstract— A Stark effect is observed in the rotational fine structure of the 0–0 band of the lowest π*π state of indole. This yields a dipole moment change |ΔμA| = 0·14±0·05. D. Evidence is presented which indicates that the sign of ΔμA is positive. An increase of dipole moment on excitation is in agreement with the value (+ 0·06 D) corresponding to the previously calculated orbitals of Song and Kurtin. The direction of the transition moment, calculated from the results of rotational analysis published by Mani and Lombardi, is found to be 26° from the molecular A-axis.  相似文献   

20.
Abstract

Errors in the calculations of Cheung and Emsley (Liquid Crystals, 13, 265 (1993)) used to derive the populations of α,δ-bis(phenoxy)ethane dissolved in isotropic and nematic liquid crystalline solvents have been corrected, and the results are compared with those obtained on the same molecule by Abe, Iizumi and Kimura (Liquid Crystals, 16, 655 (1994)). The principal difference between these two calculations is that Abe et al., obtain 0·80 as the population of the tg ± t conformers, and they find this to be independent of temperature, whereas the present calculations obtain 0·42 to 0·58 depending upon solvent and temperature.  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