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.
Peng Y  He Q  Rohani S  Jenkins H 《Chirality》2012,24(5):349-355
During the resolution of 2-chloromandelic acid with (R)-(+)-N-benzyl-1-phenylethylamine, the crystals of the less soluble salt were grown, and their structure were determined and presented. The chiral discrimination mechanism was investigated by examining the weak intermolecular interactions (such as hydrogen bond, CH/π, and van der Waals interactions) and molecular packing mode in crystal structure of the less soluble diastereomeric salt. A one-dimensional double-chain hydrogen-bonding network and a "lock-and-key" supramolecular packing mode are disclosed. The investigation demonstrates that hydrophobic layers with corrugated surfaces can fit into the grooves of one another to realize a compact packing, when the molecular structure of resolving agent is much larger than that of the racemate. This "lock-and-key" assembly is recognized to be another characteristic of molecular packing contributing to the chiral discrimination, in addition to the well-known sandwich-like packing by hydrophobic layers with planar boundary surfaces.  相似文献   

2.
3.
The optical resolution of p‐chloromandelic acid using (R)‐α‐phenylethylamine as resolving agent was presented. The effect of solvents, molar ratio of racemate to the resolving agent, filtration temperature as well as the amount of solvent on resolution was investigated by orthogonal experimentation. The binary melting point phase diagram and crystal structure analysis of diastereomeric salts rationalized the success of the resolution. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

4.
Resolution of rac‐3,3,3‐trifluorolactic acid by diastereomeric salt formation was reinvestigated. The use of (S)‐1‐phenylethylamine gives coprecipitation of two diastereomeric phases, 1 (S)‐[NH3CH(CH3)Ph](S)‐[CF3CH(OH)COO] and 2 (S)‐[NH3CH(CH3)Ph](R)‐[CF3CH(OH)COO]·H2O. Pure phase 1 may be obtained using molecular sieves as desiccants. Resolution by (S,S)‐2‐amino‐1‐phenylpropan‐1,3‐diol gives monoclinic (S,S)‐[NH3CH(CH2OH)CHOHPh] (R)‐[CF3CH(OH)‐COO] 3 with minor (S)‐3,3,3‐trifluorolactate contamination, which is precluded in the recrystallized orthorhombic form 4 . A new resolution using inexpensive phenylglycinol gives pure phase 5 (S)‐[NH3CH(CH2OH)Ph] (S)‐[CF3CH(OH)COO] in 76% yield, 94% ee in a single step, in preference to its (S)‐(R) diastereomer 6 . Overall efficient resolution for both enantiomers of the trifluorolactic acid (each ca. 70% yield, 99% ee) may be achieved by various two‐step “tandem” crystallizations, involving direct addition of either water or a second base to the filtrate from the initial reaction.  相似文献   

5.
Tang K  Yi J  Huang K  Zhang G 《Chirality》2009,21(3):390-395
This article reports a new chiral separation method-biphasic recognition chiral extraction for the separation of mandelic acid enantiomers. Distribution behavior of mandelic acid enantiomers was studied in the extraction system with O,O'-di-benzoyl-(2S,3S)-4-toluoyl-tartaric acid (D-(+)-DTTA) in organic phase and beta-CD derivatives in aqueous phase, and the influence of the types and concentrations of extractants and pH on extraction efficiency was investigated. Hydroxypropyl-beta-cyclodextrin (HP-beta-CD), hydroxyethyl-beta-cyclodextrin (HE-beta-CD), and methyl-beta-cyclodextrin (Me-beta-CD) have stronger recognition abilities for S-mandelic acid than those for R-mandelic acid, among which HP-beta-CD has the strongest ability. D-(+)-DTTA preferentially recognizes R-mandelic acid. pH and the concentrations of extractants have great effects on chiral separation ability. A high enantioseparation efficiency with a maximum enantioselectivity of 1.527 is obtained at pH of 2.7 and the ratio of 2:1 of [D-(+)-DTTA] to [HP-beta-CD]. The obtained results indicate that the biphasic recognition chiral extraction is of stronger chiral separation ability than the monophasic recognition chiral extraction. It may be very helpful to optimize the extraction systems and realize the large-scale production of pure enantiomers.  相似文献   

6.
Im SH  Ryoo JJ  Lee KP  Choi SH  Jeong YH  Jung YS  Hyun MH 《Chirality》2002,14(4):329-333
Recently, it was reported that the chiral recognition ability of (R)-N-3,5-dinitrobenzoyl phenylglycinol derivative was examined as a new HPLC chiral stationary phase (CSP 1) for the resolution of racemic N-acylnaphthylalkylamines. However, the mechanism of chiral discrimination on the CSP remained elusive until now. In this study, a spectroscopic investigation of the chiral discrimination mechanism of CSP 1 was undertaken using mixtures of (R)-N-3,5-dinitrobenzoyl phenylglycinol-derived chiral selector (2) and each of the enantiomers of N-acylnaphthylalkylamines (3) by NMR study. First, the differences in free energy changes (DeltaDeltaG) upon diastereomeric complexation in solution between the complex of each isomer with chiral selector 2 by NMR titration were calculated. The values were then compared with those estimated by chiral HPLC. The chemical shift changes of each proton on the chiral selector and analytes were also checked and it was found that the chemical shift changes decreased continuously as the acyl group on analytes increased in length. This observation was consistent with the HPLC data. From these experimental results, the interaction mechanism of chiral discrimination between the chiral selector and the analytes is more precisely explained.  相似文献   

7.
As a new acidic selector (resolving agent), we synthesized an enantiopure O‐alkyl phenylphosphonothioic acid with a seven‐membered ring ((R)‐ 5 ), which was designed on the basis of the results for the enantioseparation of 1‐arylethylamine derivatives with acyclic O‐ethyl phenylphosphonothioic acid ( I ). The phosphonothioic acid (R)‐ 5 showed unique chirality‐recognition ability in the enantioseparation of 1‐naphthylethylamine derivatives, aliphatic secondary amines, and amino alcohols; the ability was complementary to that of I . The X‐ray crystallographic analyses of the less‐ and more‐soluble diastereomeric salts showed that hydrogen‐bonding networks in the salt crystals are 21‐column‐type with a single exception which is cluster‐type. In the cases of the 21‐column‐type crystals, stability of the crystals is firstly governed by hydrogen bonds to form a 21‐column and secondly determined by intra‐columnar T‐shaped CH/π interaction(s), intra‐columnar hydrogen bond(s), inter‐columnar van der Waals interaction and/or inter‐columnar T‐shaped CH/π interaction(s). In contrast, the cluster‐type salt crystal is stabilized by the assistance of inter‐cluster T‐shaped CH/π and van der Waals interactions. To realize still more numbers of intra‐ and inter‐columnar and ‐cluster T‐shaped CH/π interactions, the seven‐membered ring of (R)‐ 5 plays a considerable role. Chirality 23:438–448, 2011. © 2009 Wiley‐Liss, Inc.  相似文献   

8.
Molecular dynamics simulations were performed on complexes of (S)-methyl N-(2-naphthyl)alaninate (NAP) with the enantiomers of N-(3,5-dinitrobenzoyl)leucine n-propylamide (DNB), which are used as models for chiral stationary-phase systems developed by Pirkle and co-workers. These studies were undertaken to qualitatively examine (pictorially) the role of entropic effects in these systems. The results of the dynamics calculations were used to refine the search for low-energy conformers. The structures were refined by the use of BioDesign's molecular mechanics method implemented in Biograf. The results of the structural refinements support our previous observation that the SR complex can achieve the same three primary interactions which are observed in the SS structure (i.e., two intermolecular hydrogen bonds and pi stacking) without a significant increase in energy. In addition, these primary interactions are conserved during molecular dynamics simulations with the occurrence of conformations which differ only in the rotational states of the alkyl side chains and ester group (which bears two potential hydrogen bond acceptors utilized in both the homo- and heterochiral complexes). The major difference in the two complexes is the relative position of the sec-butyl group and hydrogen atom on DNB's chiral center, both of which are outside the primary interaction region. All other local minima which have different relative pi orientations (“front–back,” “back–back,” and “back–front” as defined herein) are not sufficiently populated to make more than a negligible contribution to the statistical (time- or energy-averaged) analysis of the (SS)- and (SR)-NAP–DNB complexes. Thus the entropic effects observed in this study (e.g., alkyl side chain or ester group rotations) do not show evidence of qualitative differential effects on the maintenance of the same three primary interactions by both the homo- and heterochiral complexes. The reliability of the present study, which provides pictorial representations of the entropic effects, is not sufficient to determine whether the entropic effects observed herein are sufficient to achieve enantiomeric discrimination alone or in conjunction with other factors (e.g., conformational strain energy). Thus, all of the computational studies we have performed to date (i.e., our previous studies, which include strain energy and through-space field effects, and the present study, which includes entropic effects) show no evidence of any qualitative difference in the homo- and heterochiral complexes in terms of maintaining the same three “contact points”.  相似文献   

9.
Umezawa Y  Nishio M 《Biopolymers》2005,79(5):248-258
The crystal structure of acetylcholine esterase (AchE) in complex with various inhibitors, investigated as drugs for improvement of the cognitive ability of early stage Alzheimer's disease, has been analyzed with the use of our program CHPI. A number of CH/pi hydrogen bonds have been disclosed in the binding of the inhibitors with Torpedo californica AchE. It has been demonstrated that, in order to be effective in the binding with AchE, C-H bonds in the inhibitor need not be polarized.  相似文献   

10.
Coating cellulose tris (3,5‐dimethylphenylcarbamate) (CDMPC) on silica gels with large pores have been demonstrated as an efficient way for the preparation of chiral stationary phase (CSP) for high‐performance liquid chromatography (HPLC). During the process, a number of parameters, including the type of coating solvent, amount of coating, and the method for subsequent solvent removing, have been proved to affect the performance of the resultant CSPs. Coating times and the concentration of coating solution, however, also makes a difference to CSPs' performance by changing the arrangement of cellulose derivatives while remaining the coating amount constant, have much less been studied before, and thereby, were systematically investigated in this work. Results showed that CSPs with more coating times exhibited higher chiral recognition and column efficiency, suggesting that resolution was determined by column efficiency herein. Afterwards, we also investigated the effect of coating amount on the performance of CSPs, and it was shown that the ability of enantio‐recognition did not increase all the time as the coating amount; and four of seven racemates achieved best resolution when the coating amount reached to 18.37%. At the end, the reproducibility of CDMPC‐coated CSPs were further confirmed by two methods, ie, reprepared the CSP‐0.15‐3 and reevaluated the effect of coating times.  相似文献   

11.
The enantioseparation of 14 structurally similar chiral solutes, with one or two chiral centers, are studied for a commercially important polysaccharide‐based chiral stationary phase, amylose tris(3,5‐dimethylphenylcarbamate) (ADMPC). Among these solutes, only two solutes show significant enantioresolutions of 2 to 2.5 in n‐hexane/2‐propanol (90/10, v/v) at 298 K. The retention factors of the chiral solutes vary significantly from 0.7 to 7.0, and they are compared with those of simpler nonchiral solutes having similar but fewer functional groups. The sorbent–solute H‐bonding interactions between the solute functional groups and the polymer C?O and NH functional groups are probed with attenuated total reflection infrared spectroscopy (ATR‐IR). The H‐bonding interactions of the polymer C?O and NH groups with the solutes result in changes in the IR amide band wavenumbers of ADMPC upon solute adsorption. The nanostructure of an ADMPC cavity and the potential interactions with the chiral solutes are proposed based on the sorbent–solute–solvent HPLC data, the sorbent–solute IR data, and the sorbent–solute molecular dynamics (MD) simulations. The results are consistent with the three point attachment hypothesis and indicate that a significant enantioresolution in ADMPC requires at least three different interaction sites for simultaneous H‐bonds and phenyl–phenyl interactions for phenylpropylamine (PPA) and various structurally similar chiral solutes. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

12.
N-Acetyl-(E)-dehydrophenylalanine N'-methylamide [Ac-(E)-DeltaPhe-NHMe], one of a few representative (E)-alpha,beta-dehydroamino acids, was studied by FTIR in dichloromethane and acetonitrile. To support spectroscopic interpretations and to gain some deeper insight into the Ac-(E)-DeltaPhe-NHMe molecule, the Ramachandran potential energy surface was calculated by the B3LYP/6-31G*//HF/3-21G method and the conformers localized were fully optimized at the B3LYP/6-31 + G** level. The spectra and calculations were compared with those of the related molecules Ac-DeltaAla-NHMe and Ac-(Z)-DeltaPhe-NHMe. The title compound assumes two conformational states in equilibrium in dichloromethane solution with a predominance of the extended conformer E. The Ac-(E)-DeltaPhe-NHMe spectrum is like that of Ac-DeltaAla-NHMe, particularly in the region of bands AI and AII, and unlike that of Ac-(Z)-DeltaPhe-NHMe. The positions of bands AI and II together with the nu(s)(N1--H1) band proves that the conformers E of both DeltaAla and (E)-DeltaPhe compounds are stabilized by the quite strong C5 hydrogen bonds N1--H1...O2. The same conclusion is drawn from the Ramachandran diagrams. The conformers E of both compounds are placed in the global minima and the gaps in energy order between them and the second conformer are large. The conformers E of DeltaAla and (E)-DeltaPhe, apart from the N1--H1...O2 hydrogen bond, show the Cbeta--H...O1 interaction, and Ac-(E)-DeltaPhe-NHMe displays the NH/pi interaction with the N2--H2 projecting in the first carbon atom of the phenyl ring. The C5 hydrogen bond is stronger in (E)-DeltaPhe than that in the DeltaAla compound. This is in agreement with interactions found in the calculated structures and can be explained by the influence of the phenyl ring in position (E). In acetonitrile, the molecule of Ac-(E)-DeltaPhe-NHMe loses its C5 hydrogen bond and becomes unfolded, whereas that of Ac-DeltaAla-NHMe does not vary practically. Adopting conformation E in a non-polar solvent seems to be a general feature of the (E)-DeltaXaa residues.  相似文献   

13.
The FTIR spectra were analysed in the region of the nu(s)(N-H), AI(C=O) and nu(s)(Calpha=Cbeta) bands for a series of Ac-DeltaXaa-NMe2, where DeltaXaa = DeltaAla, (Z)-DeltaAbu, (Z)-DeltaLeu, (Z)-DeltaPhe and DeltaVal, to determine a predominant solution conformation of these alpha,beta-dehydropeptide-related molecules. Measurements were taken in CCl4, DCM and MeCN solutions. In the same way, spectra of saturated analogues Ac-Xaa-NMe2, where Xaa = Ala, Abu, Leu, Phe and Val, were investigated. To help interpret the spectroscopic results, conformational maps were calculated by the B3LYP/6-31+G** method. Also, the relative energies of all conformers of the dehydro compounds in vacuo as well as in the studied solvents in addition to the theoretical IR frequencies of these conformers were calculated. For comparison, molecules of two saturated analogues, Ac-L-Ala-NMe2 and Ac-L-Phe-NMe2, were calculated in a similar way. Both unsaturated and saturated compounds, which have an aliphatic side chain, occur in CCl4 and DCM mainly as a mixture of extended conformers with the C5 H-bond and open conformers. As solvent polarity increases, participation of the open conformers also increases, and in MeCN, the model amides are almost exclusively in the open form, except Ac-DeltaAla-NMe2, which shows a small amount of the H-bonded conformer. Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2 have stronger C5 hydrogen bonds than those of their saturated counterparts. As the calculations indicate, the open conformation of the unsaturated amides is conformer H/F with phi, psi -44 +/- 5 degrees, 127 +/- 4 degrees. This is the second lowest in energy conformer in vacuo and in CCl4 and the lowest one in more polar solvents. The open conformation of Ac-L-Ala-NMe2 constitutes conformer C with phi, psi -101.5 degrees, 112.7 degrees. For Ac-DeltaAla-NMe2 and Ac-DeltaAbu-NMe2, FTIR also reveals the presence of a third conformer. Calculations indicate that is the semiextended conformer D with the N1-H1...N2 hydrogen bond/contact. In all solvents, Ac-L-Phe-NMe2 and Ac-(Z)-DeltaPhe-NMe2 show only the extended E and the open H/F, respectively. In both there is an amide/pi(Ph) interaction.  相似文献   

14.
Hyun MH  Cho YJ  Song Y  Choi HJ  Kang BS 《Chirality》2007,19(1):74-81
A new doubly tethered chiral stationary phase (CSP 5) based on (+)-(18-crown-6)-2,3,11,12-tetracarboxylic acid was developed by attaching the second tethering group to silica gel through a carbon atom of the first tethering group of the corresponding singly tethered CSP (CSP 2) containing an N-CH3 tertiary amide linkage, which was previously developed in our laboratory, in order to enhance the CSP stability without the loss of chiral recognition efficiency. The new CSP was quite effective in the resolution of various racemic alpha-amino acids, amines, and amino alcohols, and the chiral recognition efficiency of the new CSP was even greater than that of the corresponding singly tethered CSP especially in terms of the resolution factors (RS). The stability of the new CSP was greater than that of the corresponding singly tethered CSP. The chromatographic resolution behaviors of the new CSP were generally consistent with those of the corresponding singly tethered CSP.  相似文献   

15.
Cellulose tris(3,5-dimethylphenylcarbamate) (CDMPC) was coated on large-pore silica gels and used as a chiral stationary phase (CSP) for high-performance liquid chromatographic separation of enantiomers. The influences of pore size of silica gel, coating amount of CDMPC, coating solvent, and column temperature on chiral discrimination were investigated. CSPs prepared with a large-pore silica gel having a small surface area showed higher chiral recognition. The amount of CDMPC adsorbed on the silica gel influenced the chiral recognition of some racemates. Loading capacity of racemates increased with an increase of the amount of CDMPC supported on the silica gel, and a CSP coated with 45% CDMPC by weight can be used for both analytical and semi-preparative scale separations. The CDMPC, coated using acetone as the coating solvent, exhibited, in many cases, higher enantioselectivity than that obtained with tetrahydrofuran F as the coating solvent. © 1996 Wiley-Liss, Inc.  相似文献   

16.
The crystal structure of class I major histocompatibility complex antigens (MHC) bound to their specific ligand peptides were analyzed, in the context of the CH/π interaction, with use of a computer program CHPI. A number of short CH/Csp2 distances have been shown at the boundary of the heavy chain and β2 microglobulin. These interactions are conserved between species, human versus murine. A number of contacts shorter than the conventional van der Waals distance have been disclosed between CH hydrogens and aromatic side-chain groups in the MHC/peptide complexes. The CH/π interaction has been suggested to contribute to the specificity in the complex formation of class I MHC.  相似文献   

17.
Kasai Y  Watanabe M  Harada N 《Chirality》2003,15(4):295-299
A convenient method for determining the absolute configuration of chiral secondary alcohols using the racemic NMR anisotropy reagent, (+/-)-2-methoxy-2-(1-naphthyl)propionic acid [(+/-)-M(alpha)NP acid], and an HPLC-CD detector was developed. The method was successfully applied to some chiral alcohols derived from (-)-alpha-santonin.  相似文献   

18.
Reaction of rac-warfarin, (?)-(1S,2R,4R)-endo-1,4,5,6,7,7-hexachlorobicyclo[2.2.1]hept-5-ene-2- carboxylic acid [(?)-HCA] and carbodiimide reagents gave two noncyclic ketonic diastereoisomeric derivatives whereas rac-warfarin and (?)-HCA acid chloride with 4-(dimethylamino)pyridine gave four cyclic hemiketal diastereoisomeric ester derivatives. The structure and stereochemistry of diastereoisomeric esters prepared from warfarin and p-chlorowarfarin were determined from 1H- and 13C-NMR spectra, mass spectra, and hydrolysis to warfarin and p-chlorowarfarin enantiomers. The structure and stereochemistry of one of the cyclic hemiketal diastereoisomeric derivatives of warfarin are supported by an X-ray crystallographic determination. Mechanisms for the formation of all products are proposed. © 1994 Wiley-Liss, Inc.  相似文献   

19.
Undecanoyl bound 3,5-dinitrobenzoyl-(S,R)-1,2-diphenylethane-1,2-diamine [(1S,2R)-DNB-DPEDA] as chiral selector (SO) has been synthesized and used as a chiral stationary phase (CSP II) for normal-phase enantioselective HPLC. It is compared with the already published diastereomeric (1S,2S)-DNB-DPEDA-derived CSP I and with the “standard” Pirkle DNB-(R)-phenylglycine-derived CSP III. Chromatographic data for about 100 racemic analytes reveal that CSP II is able to separate especially well enantiomers of derivatized aromatic carboxylic acids and analytes having a benzyl substituent bound at the chiral center. However, CSP I was found to be superior to CSP II and III in its general applicability and its ability to resolve enantiomers of heterocyclic drugs. © 1994 Wiley-Liss, Inc.  相似文献   

20.
The chiral recognition mechanism of a cinchona alkaloid-based chiral stationary phase (CSP) showing high enantiomer discrimination potential for 2-methoxy-2-(1-naphthyl)propionic acid (MalphaNP acid) was investigated. Conformational and structural analyses of the 1:1 complexes of 9-O-(tert-butylcarbamoyl) quinine selector (SO) and MalphaNP acid (selectand, SA) were carried out employing NMR spectroscopy in solution, Fourier-transform infrared (FT-IR) spectroscopy, and solid-state X-ray diffraction analysis. Intramolecular NOEs of a soluble analogue of the CSP afforded the conformational states of the free and complexed form of the selector. The (1)H-NMR spectra revealed that the free form of the SO constitutes anti-open as well as anti-closed and/or syn-closed conformers. Upon complexation with the (S)-MalphaNP acid enantiomer to form the more stable diastereomeric associate, a conformational transition of the selector takes place, resulting in the synthesis of the anti-open conformer nearly exclusively. FT-IR spectra reveal that, besides the primary ion-pairing interaction, stereoselective hydrogen bonding stabilizes the more stable complex via the amide hydrogen of the SO. X-ray diffraction analysis of 9-O-(tert-butylcarbamoyl)quinine and (S)-MalphaNP acid complex further revealed the occurrence of a bidentate H-bond-mediated ionic interaction between SO and SA as well as the lack of pi-pi interaction in the 1:1 complex, and corroborated the conclusions derived from spectroscopic and chromatographic studies.  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