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 78 毫秒
1.
Scalemic mixtures of chiral anisyl fenchols with different ortho-substituents (X) in the anisyl moieties [X = H (1), Me (2), SiMe3 (3) and tBu (4)] are employed as pre-catalysts in enantioselective additions of diethylzinc to benzaldehyde. While a remarkable asymmetric depletion is apparent for X = H and Me, a linear relationship between the enantiomeric purity of the chiral source and the product 1-phenylpropanol is observed for X = SiMe3 and tBu. X-ray single crystal analyses show that racemic methylzinc fencholates obtained from 1 (X = H) and 2 (X = Me) yield homochiral dimeric complexes, while for 3 (X = SiMe3) and 4 (X = tBu) the heterochiral dimeric alkylzinc structures are formed. The enantiopure fenchols 1-4 all yield homochiral dimeric methylzinc complexes. Computed relative energies of homo- and heterochiral fencholate dimers with X = H and Me reveal an intrinsic preference for the formation of the homochiral dimers, consistent with the observed negative NLE. In contrast, similar stabilities are computed for homo- and heterochiral complexes of ligands 3 (X = SiMe3) and 4 (X = tBu), in agreement with the absence of a nonlinear effect for bulky ortho-subsituents.  相似文献   

2.
The chiral discrimination in the self-association of chiral 1,3a,4,6a-tetrahydroimidazo[4,5-d]imidazoles has been studied using density functional theory methods. Clusters from dimers to heptamers have been considered. The heterochiral dimers (RR:SS or SS:RR) are more stable than the homochiral ones (RR:RR or SS:SS) with energy differences up to 17.5 kJ/mol. Besides, in larger clusters the presence of two adjacent homochiral molecules impose an energetic penalty when compared to alternated chiral systems (RR:SS:RR:SS...). The differences in interaction energy within the dimers of the different derivatives have been analyzed based on the atomic energy partition carried out within the atoms in molecules framework. The mechanism of proton transfer in the homo- and heterochiral dimers shows large transition-state barriers except in those cases in which a third additional molecule is involved in the transfer. The optical rotatory power of several clusters of the parent compound have been calculated and rationalized based on the number of homochiral interactions and the number of monomers of each enantiomer within the complexes.  相似文献   

3.
Unique perylene diastereomeric linear and cyclic dimers were synthesized from twisted perylene monomers, revealing that pi-stacking stereoisomerism imparted specific intermolecular self-assembly and intramolecular folding. Only the homochiral twisted tetrachloroperylene monomers cyclized via a cooperative reaction, forming the homochiral diastereomers. The heterochiral tetrachloroperylene monomers proceeded through a stepwise reaction and yielded a linear heterochiral dimer, which equilibrated with the linear homochiral dimers. The linear homochiral dimers cyclized to produce the same cyclic homochiral diastereomers. These results demonstrated that homochiral and heterochiral self-assemblies were two distinct molecular codes, directing two specific chemical pathways. The homochiral cyclic dimers remain isomerically pure at -20 degrees C but can be interconverted to the heterochiral cyclic dimer meso compound at room temperature. The diastereomers were readily separated by HPLC. While driven by solvophobic forces, foldable linear dimers synthesized from the same twisted monomers using phosphoramidite chemistry folded into homodimer and heterodimer, confirming the inherent molecular codes, which were dictated by the perylene chirality, ultimately gauged the weak pi-stack forces, and directed self-assembly and folding.  相似文献   

4.
The homochiral and heterochiral hydrogen-bonded (HB) dimers of a set of small model molecules (alpha-amino alcohols) have been studied by means of ab initio methods. The gas-phase calculations have been carried out with the hybrid HF/DFT B3LYP method and the 6-311++G** basis set. The electron density of the complexes has been analyzed using the atoms in molecules (AIM) methodology, which allows characterization of the HB interactions and additional intermolecular contacts. To take into account the water solvation effect, the polarized continuum model (PCM) method has been used to evaluate the Delta G(solv). The gas-phase results show that the heterochiral dimers are the most stable ones for each case studied, while in solution for several cases, the relative stability is reversed and the homochiral dimers become more stable. The AIM analysis shows the typical bond critical points characteristic of the HB and additional bond critical points denoting, in this case, destabilization of intermolecular interaction as CF(3)...F(3)C and CH(3)...H(3)C contacts.  相似文献   

5.
This communication asks whether homochiral or heterochiral interaction is preferred between enantiomeric beta-sheets and finds that homochiral pairing is strongly preferred. Interactions between beta-sheets occur widely among proteins through pairing of the hydrogen-bonding edges. Although the hydrogen-bonding edges of both l- and d-beta-sheets put forth the same pattern of hydrogen-bond donor and acceptor groups, the side chains point in opposite directions. Homochiral pairing of beta-sheets generates structures in which the pleats and side chains of adjacent beta-strands are parallel to each other, while heterochiral pairing of beta-sheets generates structures in which the pleats and side chains are antiparallel. To test which pairing is preferred, we have prepared and studied the interactions of beta-sheets 1a-d, which comprise all l-amino acids, and beta-sheets 2a-c, which comprise all d-amino acids. Previous studies in our laboratory have established that these compounds form well-defined dimers in organic solvents. In the current study, 1H NMR experiments establish that when the l-beta-sheets (1) are mixed with the enantiomeric d-beta-sheets (2), homochiral beta-sheet dimers predominate, and only small quantities of heterochiral beta-sheet dimers form. Ratios of homochiral and heterochiral dimers ranging from 95.8:4.2 to 98.5:1.5 are measured in CDCl3 at 253 K, which correspond to statistically corrected free-energy differences of 3.1-4.2 kcal/mol (0.6-0.8 kcal/mol per interacting residue). Possible explanations for the high enantioselectivity of molecular recognition between beta-sheets include favorable nonbonded contacts between the adjacent beta-strands of the homochiral beta-sheets and poor fit of the heterochiral beta-strands, which should twist in opposite directions.  相似文献   

6.
Stability and structure of homo- and heterochiral protonated serine (Ser) dimers were investigated by a combination of mass spectrometry and ab initio quantum chemical calculations. This established that the energy difference between the most stable homo- and heterochiral forms is very small: tandem mass spectrometry with Cooks' kinetic method yielded a negligible difference in Gibbs free energy (0.2+/-0.2 kJ mol(-1)). The various isomeric forms of (Ser)2 H+ and their energetics were determined by extensive electronic-structure calculations, which yielded homo- and heterochiral forms of the isomers with distinctly different relative energies. The most stable homochiral isomer is stabilized by two hydrogen bonds and is far more stable than any other homochiral isomer. The most stable heterochiral isomer has completely different features, and it is characterized by a salt-bridge structure. This clearly shows that salt-bridge structures do exist in the gas phase even in comparatively small molecules and in the absence of particularly basic or acidic functional groups.  相似文献   

7.
The theoretical study of the stereoisomer discrimination of the 2:1 homo- and heterochiral complexes between chiral 5,5'-dihalogen bifuranes and lithium or sodium cations has been carried out using DFT methods. To understand the chiral effect produced by the introduction of a second bifurane molecule, the 1:1 complexes also have been calculated. All the 2:1 heterochiral complexes computed showed a nonplanar configuration around the metallic cation, but in the case of the homochiral complexes, the dibromo- and dichloro-bifurane systems around Na+ were quasi-planar. The nature of the interactions established between cations and bifurane systems has been analyzed by means of AIM and NBO, and correlations between the electron density topological parameters with the O...M distance and with the orbital interaction energy have been found. Stereodiscrimination is observed favoring the heterochiral complexes except for the Na+ complexes with chloro and bromo substituents in which the homochiral forms are more stable. Stereodiscrimination values correlate with the difference in electron density at the bond critical point and orbital interaction energy between homo- and heterochiral systems.  相似文献   

8.
Chiral recognition has been studied in neutral or ionic weakly bound complexes isolated in the gas phase by combining laser spectroscopy and quantum chemical calculations. Neutral complexes of the two enantiomers of lactic ester derivatives with chiral chromophores have been formed in a supersonic expansion. Their structure has been elucidated by means of IR-UV double resonance spectroscopy in the 3 μm region. In both systems described here, the main interaction ensuring the cohesion of the complex is a strong hydrogen bond between the chromophore and methyl-lactate. However, an additional hydrogen bond of much weaker strength plays a discriminative role between the two enantiomers. For example, the 1:1 heterochiral complex between R-(+)-2-naphthyl-ethanol and S-(+) methyl-lactate is observed, in contrast with the 1:1 homochiral complex which lacks this additional hydrogen bond. On the other hand, the same kind of insertion structures is formed for the complex between S-(±)-cis-1-amino-indan-2-ol and the two enantiomers of methyl-lactate, but an additional addition complex is formed for R-methyl-lactate only. This selectivity rests on the formation of a weak CHπ interaction which is not possible for the other enantiomer. The protonated dimers of Cinchona alkaloids, namely quinine, quinidine, cinchonine and cinchonidine, have been isolated in an ion trap and studied by IRMPD spectroscopy in the region of the ν(OH) and ν(NH) stretch modes. The protonation site is located on the alkaloid nitrogen which acts as a strong hydrogen bond donor in all the dimers studied. While the nature of the intermolecular hydrogen bond is similar in the homochiral and heterochiral complexes, the heterochiral complex displays an additional weak CHO hydrogen bond located on its neutral part, which results in slightly different spectroscopic fingerprints in the ν(OH) stretch region. This first spectroscopic evidence of chiral recognition in protonated dimers opens the way to the study of the complexes of Cinchona alkaloids involved in enantioselective catalysis. These examples show how secondary hydrogen bonds controlled by stereochemical factors govern molecular recognition processes.  相似文献   

9.
The mechanism of the Soai reaction has been thoroughly investigated at the M05‐2X/6‐31G(d) level of theory, by considering ten energetically distinct paths. The study indicates the fully enantioselective catalytic cycle of the homochiral dimers to be the dominant mechanism. Two other catalytic cycles are shown to both be important for correct understanding of the Soai reaction. These are the catalytic cycle of the heterochiral dimer and the non‐enantioselective catalytic cycle of the homochiral dimers. The former has been proved to be not really competitive with the principal cycle, as required for the Soai reaction to manifest chiral amplification, whereas the latter, which is only slightly competitive with the principal one, nicely explains the experimental enantioselectivity observed in the reaction of 2‐methylpyrimidine‐5‐carbaldehyde. The study has also evidenced the inadequacy of the B3LYP functional for mechanistic investigations of the Soai reaction.  相似文献   

10.
Diastereomeric adducts between (S)-1-(4-fluorophenyl)-ethanol and R and S 2-butanol, formed by supersonic expansion, have been investigated by means of a combination of mass selected resonant two-photon ionization-spectroscopy and infrared depletion spectroscopy. Chiral recognition is evidenced by the specific spectroscopic signatures of the S(1)← S(0) electronic transition as well as different frequencies and intensities of the OH stretch vibrational mode in the ground state. D-DFT calculations have been performed to assist in the analysis of the spectra and the determination of the structures. The homochiral and heterochiral complexes show slight structural differences, in particular in the interaction of the alkyl groups of 2-butanol with the aromatic ring. The experimental results show that the homochiral [FE(S)·B(S)] complex is more stable than the heterochiral [FE(S)·B(R)] diastereomer in both the ground and excited states. The binding energy difference has been evaluated to be greater than 0.60 kcal mol(-1).  相似文献   

11.
The chiral recognition in the formation of hydrogen-bonded (HB) dimers of 1,6a-dihydropyrrolo[2,3-b]pyrrole derivatives as well as in their proton-transfer processes have been studied by means of ab initio calculations. The heterochiral dimers are in general the most stable ones, but amphiprotic substituents that are able to form attactive interactions with twin groups revert this tendency. Energy differences up to 4.0 kcal/mol have been found favoring the homo- or heterochiral complexes. Two possible proton-transfer processes have been studied, the concerted one and the nonconcerted one. The compresion of the systems in the transition structures produce an increase in the energetic differences when compared to the corresponding minima complexes. A Steiner-Limbach relationship has been found for the geometrical properties of the HB in the minima and transition states calculated here. The electron density and its Laplacian at the bond critical point have been found to correlate with the HB distance.  相似文献   

12.
Chiral recognition and subsequent selective self-organisation into hydrogen-bonded n-mers is observed in supersonic methyl lactate expansions. The nu(OH) and nu(C=O)-vibrations are investigated by ragout-jet FTIR-spectroscopy and lead to the assignment of homo- and heterochiral clusters of at least three different cluster sizes. Whereas homo- and heterochiral dimers are formed in similar amounts in the racemic mixture, prominent absorptions due to different homochiral and heterochiral lactate trimers and tetramers indicate highly specific chiral self-recognition beyond molecular pairs. Chemical modification of the ester-group (methyl-, ethyl- and isopropyl-lactate) and argon admixture to the helium expansion contribute importantly to an understanding of the cluster spectra and topology.  相似文献   

13.
The study of possible chiral recognition of a series of peptide models (For-Gly-NH2, For-Ala-NH2 and four of their fluoro substituted derivatives) has been carried out by means of DFT calculations. Homo (L,L) and heterochiral (L,D) dimers formed by hydrogen bond (HB) complexation have been considered. Initially, the conformational preferences of the monomers have been calculated and used to generate all the possible homo and heterochiral dimers. The energetic results show that in most cases, the β monomers are the most stable while in the dimers, the γ–γ complexes show the strongest interaction energies. In three of the four chiral cases studied, a heterochiral dimer is the most stable one. In addition, the electron density and nuclear shielding of the complexes have been studied.  相似文献   

14.
A theoretical study of the dimer formation of chiral 1,8a-dihydro-1,8-naphthyridine derivatives has been carried out by means of DFT calculations. In the cases treated, the heterochiral dimers (RS or SR) are always more stable than the homochiral ones (RR or SS). Two possible proton transfer processes have been studied, the concerted and the non-concerted ones. The non-concerted TS corresponds to a true TS while the concerted one presents two imaginary frequencies. The geometrical characteristics of the hydrogen bonds in all the structures calculated have been correlated using the Steiner–Limbach model.  相似文献   

15.
16.
In the present work, we compare the intermolecular energy surfaces of the alanine molecule in its neutral and zwitterionic state using ab initio theory (HF/6-311++G) as a function of mutual orientation. Starting from the optimized structures of the nonbonded homochiral (l-l) and heterochiral (d-l) pairs of molecules, the energy surfaces are studied with rigid geometry by varying the distance and orientation. The potential energy surfaces of the l-l and d-l pairs are found to be dissimilar and reflect the underlying chirality of the homochiral pair and racemic nature of the heterochiral pair. The intermolecular energy surface of the l-l pair is more favorable than the corresponding energy surface of the d-l pair. The study, for the first time, reveals clear homochiral preference without use of parameters, which was unobserved in previous detailed simulations but predicted by theory. The electrostatic interaction further augments the chiral discrimination. The basis set superposition error (BSSE) corrected results show enhanced discrimination. Use of higher-level M?ller-Plesset perturbation theory (MP2) and further BSSE correction do not change the conclusions made at the Hartree-Fock (HF) level. The major conclusions based on HF and MP2 level calculations remain unaltered when the calculations of the potential energy surfaces for the neutral and zwitterionic pairs are repeated using the density functional theory (DFT) (B3LYP/6-311++G). The observed orientation dependence has significance in the biological chiral recognition as well as peptide synthesis at the peptidyl transferase center where the amino terminal and peptidyl terminal undergo mutual rotatory motion.  相似文献   

17.
The phenomenon of chiral crystallization into homochiral crystals is known for more than 170 years, yet it is still poorly understood. Studying crystallization on surfaces under well-defined condition seems a promising approach towards better understanding the intermolecular chiral recognition mechanisms during nucleation and growth. The two-dimensional aggregation of racemic trioxaundecahelicene on the single crystalline silver(100) surface has been investigated with scanning tunneling microscopy and with non-contact atomic force microscopy, as well as molecular modeling simulations. A transition from homochiral cluster motifs to heterochiral assembly into large islands with increasing coverage is observed. Force field modelling confirms higher stability of heterochiral arrangements from twelve molecules on. Results are discussed with respect to previous findings for the all-carbon heptahelicene on the same surface.  相似文献   

18.
In this work, we investigate the mode of chiral interactions between the asymmetric Co(II) salen complex, (S,S)-N,N'-bis(3,5-di-tert-butylsalicylidene)-1,2-cyclohexane-diamine-Co(II) ([Co(1)]), and single enantiomers of methylbenzylamine (MBA) using different continuous-wave and pulsed electron paramagnetic resonance techniques combined with density functional theory computations. While [Co(1)] displays a large affinity for binding a single MBA molecule, it has a much weaker affinity for binding a second MBA molecule. Subtle differences are detected in the EPR spectra of the homochiral (S,S-[Co(1)](S-MBA)) and heterochiral (S,S-[Co(1)](R-MBA)) adducts using low [Co(1)] : MBA ratios. Moreover at high concentrations of racemic MBA, a strong preference (80%) is observed for the formation of the bis-ligated heterochiral adduct (S,S-[Co(1)](R-MBA)(2)) compared to the homochiral analogue (20% of S,S-[Co(1)](S-MBA)(2)). Differences in the (14)N hyperfine coupling from the diamine backbone in [Co(1)] were also evidenced by hyperfine sublevel correlation (HYSCORE), revealing magnetically equivalent N nuclei for the homochiral adducts and inequivalent N nuclei for the heterochiral adducts. Using DFT, these slight differences were reproduced, and explained based upon the different modes of alignment of the MBA molecule in the adduct. The current findings therefore reveal the appreciable enantiodiscrimination that occurs during the binding of MBA enantiomers to the chiral Co(II) salen complex.  相似文献   

19.
Directing the supramolecular polymerization towards a preferred type of organization is extremely important in the design of functional soft materials. Proposed herein is a simple methodology to tune the length and optical chirality of supramolecular polymers formed from a chiral bichromophoric binaphthalene by the control of enantiomeric excess (ee). The enantiopure compound gave thin fibers longer than a few microns, while the racemic mixture favored the formation of nanoparticles. The thermodynamic study unveils that the heterochiral assembly gets preference over the homochiral assembly. The stronger heterochiral binding over homochiral one terminated the elongation of fibrous assembly, thus leading to a control over the length of fibers in the nonracemic mixtures. The supramolecular polymerization driven by π–π interactions highlights the effect of the geometry of a twisted π‐core on this self‐sorting assembly.  相似文献   

20.
A theoretical study of the chiral distinction between the homochiral and heterochiral dimers of the 1-aza[n]helicenes, with n = 1–7, glued with lithium, sodium, and potassium cations has been carried out by means of DFT calculations up to M05-2x/6-311+G(d) computational level. The electronic characteristic of the isolated helicenes has been explored. The chiral distinction is dependent on the size of the helicene and the cation used with the largest values obtained for the 1-aza[6]helicene bound to lithium cation.  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