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 16 毫秒
1.
The influence of a chiral surfactant and a polymer-supported chiral additive on reduction of ketones using sodium borohydride will be described. Initial preparations involved methylation of (S)-leucinol to give (2S)-N , N-dimethyl-2-amino-4-methyl-1-pentanol (1) (67%). The chiral surfactant (2) was synthesized by reacting (1) with bromohexadecane (71%). The functionalized styrene for the polymer-supported chiral additive (5) was synthesized by reacting (1) with 4-vinylbenzyl chloride. Polymerization was carried out with 10% of the functionalized monomer (4), 5% cross-linking agent divinylbenzene, and 85% styrene with AIBN as the initiator. The activity of the chiral surfactant and polymeric additive were examined by using them as additives in a standard reduction of 2-pentanone with sodium borohydride to yield (R)- and (S)-2-pentanol (3) (20%). The resulting alcohol was analyzed by polarimetry (ee 9.5%) and also esterified with (2S)-methylbutyric acid prior to characterization by NMR. 13C NMR indicated an enantiomeric excess of 5.2% when the chiral surfactant was used, and 7% when the polymeric additive was used.  相似文献   

2.
Enantiospecific adsorption of cysteine molecules onto chiral kink sites on the Au(110)-(1x2) surface was observed by scanning tunneling microscopy. l- and d-cysteine dimers were found to adopt distinctly different adsorption geometries at S kinks, which can be understood from the need to reach specific, optimum molecule-substrate interaction points. Extended, homochiral domains of l/d-cysteine were furthermore observed to grow preferentially from R/S kinks. The results constitute the first direct, microscopic observation of enantiospecific molecular interaction with chiral sites on a metal single-crystal surface.  相似文献   

3.
Thin films with enantioselective properties for electrochemically active chiral probes were developed. Enantioselectivity was accomplished via molecular imprinting. The films were fabricated through the sol-gel technique and were spin-coated on ITO electrodes. The chiral selectivity recognition was detected using two enantiomer pairs: D- and L-3,4-dihydroxyphenylalanine (D- and L-dopa) and (R)- and (S)-N,N'-dimethylferrocenylethylamine [(R)-Fc and (S)-Fc]. A defined chiral cavity was obtained by selection of functional monomers that interact with the template molecule, followed by its removal. Chiral selection properties were measured by cyclic voltammetry and square wave voltammetry. For both template molecules, very good chiral recognition was revealed by electrochemical measurement. The nonspecific adsorption measured for reference nonimprinted films was negligible (less than 5%). Dopa imprinted films revealed both high sensitivity, by the detection of 1 nM (0.2 ppb) concentration, and excellent selectivity, when challenged with a series of catechol derivatives. Fc-imprinted films were able to detect ca. 2 ppm of the target molecule, with very good enantioselectivity and low nonspecific adsorption. To our knowledge, this is the first report of successful molecular imprinting of a ferrocene derivative.  相似文献   

4.
The hemilabile chiral C2 symmetrical bidentate substituted amide ligands (1R,2R)-5(a-d) and (1S,2S)-6(a-d) were synthesized in quantitative yield from (1R,2R)-(+)-3-methylenecyclo-propane-1,2-dicarboxylic acid (1R,2R)-3 and (1S,2S)-(-)-3-methylene-cyclopropane-1,2-dicarboxylic acid (1S,2S)-3, in two steps, respectively. The chiral Feist's acids (1R,2R)-3 and (1S,2S)-3 were obtained in good isomeric purity by resolution of trans-(±)-3-methylene-cyclopropane-1,2-dicarboxylic acid from an 8:2 mixture of tert-butanol and water, using (R)-(+)-α-methylbenzyl amine as a chiral reagent. This process is reproducible on a large scale. All these new synthesized chiral ligands were characterized by 1H-NMR, 13C-NMR, IR, and mass spectrometry, as well as elemental analysis and their specific rotations were measured. These new classes of C2 symmetric chiral bisamide ligands could be of special interest in asymmetric transformations.  相似文献   

5.
(R)- or (S)-2-Methylferrocene carboxylic acids, (R)-1 or (S)-1, (R)- or (S)-2-phenylbutanoic acid, (R)-2 or (S)-2, and (R)- or (S)-2-propanoic acid, (R)-3 or (S)-3, can be imprinted in thin TiO2 films on the gate surface of ion-sensitive field-effect transistor (ISFET) devices. The imprinting is performed by hydrolyzing the respective carboxylate TiIV butoxide complex on the gate surface, followed by washing off the acid from the resulting TiO2 film. The imprinted sites reveal chiroselectivity only towards the sensing of the imprinted enantiomer. The chiral recognition sites reveal not only chiroselectivity but also chirospecificity and, for example, the (R)-2-imprinted film is active in the sensing of (R)-2, but insensitive towards the sensing of (R)2-phenylpropanoic acid, (R)-3, which exhibits a similar chirality. Similarly, the (R)-3-imprinted film is inactive in the analysis of (R)-2. The chiroselectivity and chirospecificity of the resulting imprinted films are attributed to the need to align and fit the respective substrates in precise molecular contours generated in the cross-linked TiO2 films upon the imprinting process.  相似文献   

6.
Ketoreductase (KR) activities help determine the stereochemistry of the products of modular polyketide synthases (PKSs). For example, domains eryKR(1) and eryKR(2), contained, respectively, in the first and second extension modules of the erythromycin-producing PKS, reduce 3-ketoacyl-thioester intermediates with opposite stereospecificity. Amino acid motifs that correlate with stereochemical outcome have been identified in KRs. We have used saturation mutagenesis of these motifs in eryKR(1) and eryKR(2), and a microplate-based screen of such mutants for activity against (9R, S)-trans-1-decalone, to identify candidate enzymes potentially altered in stereocontrol. Active mutants were reassayed with (2R, S)-2-methyl-3-oxopentanoic acid N-acetylcysteamine thioester, and the alcohol products were analyzed by chiral HPLC. Variant enzymes were found with either altered substrate selectivity for the (2R) or (2S) substrate or altered stereospecificity of reduction, or both, further highlighting the importance of these motifs in stereochemical control.  相似文献   

7.
Immobilized amyloglucosidase was used as a chiral stationary phase (CSP). First, the retention and enantioselectivity of several model chiral amines and acids were investigated. We found that this CSP was unable to separate the enantiomers of acids, though all selected amines could be resolved. The adsorption of (R)- and (S)-propranolol and its influence on column temperature and 2-propanol content in the eluent were then studied in detail, using a three-step methodology. The adsorption was first evaluated using Scatchard plots; thereafter, the adsorption was characterized in detail by calculating the adsorption energy distribution. With this model-independent information, a better judgment could be made of the possible adsorption models selected in the last step, the model fitting to the data. In the case examined, the bi-Langmuir model (containing nonselective and enantioselective sites) describes the system well. The retention of (R)- and (S)-propranolol at low temperatures increases with the content of 2-propanol in the eluent, due to the increased saturation capacity of the enantioselective sites. The retention is an enthalpy-driven process at both types of sites, whereas the enantioseparation is due to differences between the entropy changes of the two enantiomers at the enantioselective sites. The enthalpy of adsorption at the nonselective sites is almost identical at the two concentrations of 2-propanol in the eluent. Enantioselective adsorption, on the other hand, is more exothermic at higher modifier content (20%). Thus, at high temperatures the retention decreases with increasing modifier content, whereas the opposite (unusual) trend is the case at low temperatures.  相似文献   

8.
The adsorption rates onto a range of platinum single-crystal surfaces of key species involved in the proline-directed heterogeneous enantioselective hydrogenation of isophorone were investigated by electrochemical means. Specifically, the uptakes of the prochiral reactant (isophorone), the chiral hydrogenation product (3,3,5-trimethylcyclohexanone), and the chiral directing agent ((R)- and (S)-proline) were examined. The effects of R,S chiral kink sites on the adsorption of (R,S)-proline were also studied. The reactant adsorbs approximately 105 times faster than the chiral modifier so that under conditions of competitive adsorption the latter is entirely excluded from the metal surface. Supplementary displacement and reaction rate measurements carried out with practical Pd/carbon catalysts show that under certain reaction conditions isophorone quickly displaces preadsorbed proline from the metal surface. Thus both kinetics and thermodynamics ensure that the chiral modifier can play no role in any surface-mediated process that leads to enantiodifferentiation. These results are fully consistent with the recent proposal1 that the crucial step leading to enantiodifferentiation occurs in the solution phase and not at the metal surface. In addition, it is found that there is no preferred diastereomeric interaction between (R,S)-proline and R,S step kink sites on Pt{643} and Pt{976}, implying that such sites do not play a role in determining the catalytic behavior of supported metal nanoparticles.  相似文献   

9.
The adsorption and thermal chemistry of 2-butanol and propylene oxide, each individually and when coadsorbed together, were characterized on Pt(111) single-crystal surfaces by using temperature programmed desorption and reflection-adsorption infrared spectroscopies. The formation of chiral superstructures on the surface upon the deposition of submonolayer coverages of enantiopure 2-butoxide species, produced by thermal dehydrogenation of 2-butanol, was highlighted by their difference in behavior toward the adsorption of the two enantiomers of propylene oxide. It was found that a significant enhancement in adsorption is possible on surfaces with the same chirality of the probe molecule, that is, for (R)-propylene oxide adsorption on (R)-2-butoxide layers and for (S)-propylene oxide adsorption on (S)-2-butoxide layers. The propylene oxide probe was found to also adsorb with the ring closer to the surface in those cases. Finally, less butoxide decomposition is seen at higher temperatures from the homochiral pairing, presumably because the coadsorbed propylene oxide forces the alkoxides into a more compact and better packed structure on the surface.  相似文献   

10.
The adsorption and thermal chemistry of (S)-(+)-2-methylbutanoic acid ((S)-2MBA) on Pt(111) single-crystal surfaces was characterized by using temperature programmed desorption (TPD) and reflection-adsorption infrared (RAIRS) spectroscopies. Particular emphasis was placed on the characterization of the chiral superstructures formed upon the deposition of the submonolayer coverages of enantiopure (S)-2-methylbutanoate species that are produced by thermal dehydrogenation of the (S)-2MBA. The enantioselectivity of the empty platinum sites left open on those structures were identified by their difference in behavior toward the adsorption of the two enantiomers of propylene oxide. It was found that a significant enhancement in adsorption is possible on surfaces with the same chirality of the probe molecule, specifically that the uptake of (S)-propylene oxide is larger than that of (R)-propylene oxide on (S)-2-methylbutanoate adsorbed layers. This contrasts with the lack of enantioselectivity previously reported for the same adsorbate on Pd(111). Detectable differences in adsorption energetics of (R)- vs (S)-propylene oxide on the (S)-2-methylbutanoate/Pt(111) overlayers were measured but deemed not to be the controlling factor in the enantioselectivity reported in this system.  相似文献   

11.
Kahle KA  Foley JP 《Electrophoresis》2006,27(4):896-904
In this study, the combination of two chiral components in a microemulsion formulation for the separation of enantiomers via microemulsion EKC (MEEKC) was successfully accomplished. Previous publications of chiral microemulsions have utilized only one chiral entity; the surfactant, cosurfactant, or oil was chiral. This is the first study, to date, of the effects of using two chiral species in a single pseudostationary phase (PSP). The chiral surfactant dodecoxycarbonylvaline (DDCV) was used in conjunction with the chiral cosurfactant S-2-hexanol. Ethyl acetate was incorporated as the oil core of the microemulsion and the buffer was 50 mM phosphate at a pH of 7. Additionally, a microemulsion prepared with racemic 2-hexanol was used for comparison to a previous DDCV microemulsion and as a baseline for the newly formulated dual chiral microemulsion. The efficiencies, resolutions, and enantioselectivities for the S-2-hexanol, racemic 2-hexanol, and original 1-butanol DDCV microemulsions are compared. The hexanol-based PSPs provide improved efficiencies and resolutions. To evaluate the combination of each DDCV enantiomer (R and S) with S-2-hexanol, changes in Gibb's free energy were calculated. A synergistic effect was found when two chiral components were combined to form a microemulsion.  相似文献   

12.
An S‐mandelic acid imprinted chitosan resin was synthesized by cross‐linking chitosan with glutaraldehyde in 2% acetic acid solution. S‐Mandelic acid imprinted chitosan resin was used to enantioselectively separate racemic mandelic acid in aqueous medium. When keeping the pH of sample solution (100 mM Tris‐H3PO4) at 3.5 and adsorption time at 40 min, the enantiomer excess of mandelic acid in supernatant was 78.8%. The adsorption capacities of S‐mandelic acid imprinted chitosan resin for S‐ and R‐mandelic acid were determined to be 29.5 and 2.03 mg/g, respectively. While the adsorption capacities of non‐imprinted cross‐linked chitosan for S‐ and R‐mandelic acid were 2.10 and 2.08 mg/g, respectively. The result suggests that the imprinted caves in S‐mandelic acid imprinted chitosan resin are highly matched with S‐mandelic acid molecule in space structure and spatial arrangement of action sites. Interestingly, the enantiomer excess value of mandelic acid in supernatant after adsorption of racemic mandelic acid by R‐mandelic acid imprinted cross‐linked chitosan was 25.4%. The higher enantiomer excess value by S‐mandelic acid imprinted chitosan resin suggests that the chiral carbons in chitosan and the imprinted caves in S‐mandelic acid imprinted chitosan resin combine to play roles for the enantioselectivity of S‐mandelic acid imprinted chitosan resin toward S‐mandelic acid. Furthermore, the excellent enantioselectivity of S‐mandelic acid imprinted chitosan resin toward S‐mandelic acid demonstrates that using chiral chitosan as functional monomer to prepare molecularly imprinted polymers has great potential in enantioseparation of chiral pharmaceuticals.  相似文献   

13.
Unknown diaryl (1-3) and alkyl-phenyl (4, 5) N-tosyl aziridines have been successfully synthesized from pure (R,R,R,S(S))-(-)-sulfonium salt derived from Eliel's oxathiane, tosylimines 11a-f, and using a phosphazene base (EtP(2)) to generate the ylide. Both cis and trans aziridines have exceptionally high enantiomeric purities (98.7-99.9%). The (2R,3R)-configuration of trans-3 and the (2R,3S)-configuration of cis-4 have been determined by X-ray analysis using the Bijvoet method. The R-configuration found at C2 is consistent with the model and all previous results, therefore all trans-aziridines and cis-aziridines have been assigned the (2R,3R)- and the (2R,3S)-configurations, respectively. This two-step asymmetric synthesis can be easily used on gram quantities and involves no unstable/hazardous reagent. The chiral auxiliary is used in a stoichiometric amount but is recovered in high yield and reused.  相似文献   

14.
Enantiomerically pure (3'R)- and (3'S)-3',4'-dihydrospiro[piperidine-2,3'(2'H)-benzopyran]s (R)-10 and (S)-10 were successfully synthesized according to the CN(R,S) methodology with the aim of serving as a pattern for the generation of related spirocyclic compounds. Two different synthetic pathways were studied starting from 2-cyano-6-phenyloxazolopiperidine (-)-2. One of them was selected and used for the preparation of amines (R)-17 and (S)-17 starting from (-)-2 and (+)-2, respectively. The enantiomeric purity of all final aminochroman derivatives was determinated by capillary electrophoresis using beta-cyclodextrin as the chiral selector.  相似文献   

15.
Mass spectrometry (MS) has become a popular analytical technique because of its high sensitivity and specificity. Therefore, the use of a chiral derivatization reagent for the MS detection seems to be efficient for the enantiomeric separation of racemates. However, the number of chiral reagents for the liquid chromatography (LC)-tandem mass spectrometry (MS/MS) analysis is very limited. The applicability of commercially available chiral amines as the derivatization reagents for the enantiomeric separation of chiral carboxylic acids is reported in this paper by using non-steroidal anti-inflammatory drugs (NSAIDs), i.e. ibuprofen, flurbiprofen, and loxoprofen. The efficiency of the chiral reagents was evaluated in terms of tagging easiness, separation by reversed-phase chromatography, and detection sensitivity by electrospray ionization (ESI)-MS/MS. Among the tested eight chiral amines, i.e. (R)-(+)-4-(3-aminopyrrolidin-1-yl)-7-(N,N-dimethylaminosulfonyl)-2,1,3-benzoxadiazole (DBD-APy), (S)-(+)-1-(2-pyrrolidinylmethyl)-pyrrolidine (PMP), L-prolinamide, (3R)-(-)-1-benzyl-3-aminopyrrolidine, (S)-(+)-1-cyclohexyl-ethylamine, (3R)-(+)-3-(trifluoroacetamido)-pyrrolidine (TFAP), (R)-(-)-1-aminoindan (AI), and (S)-(+)-tetrahydrofurfuryl-amine, DBD-APy, PMP, AI, and TFAP could be used as the chiral reagents for the enantiomeric separation of the NSAIDs. The Rs values and the detection limits of the derivatives were in the range of 1.29-3.85 and 0.57-0.96 fmol, respectively. These four reagents were applied for the determination of the NSAIDs in rat plasma.  相似文献   

16.
A comprehensive study into the coordination chemistry of two C3-chiral tripodal amido ligands has been carried out. The amido ligands contain a trisilylmethane backbone and chiral peripheral substituents. The amine precursors. HC(SiMe2NH[(S)-1-phenylethyl]]3 (1) and HC[SiMe2NH[(R)-1-indanyl]]3 (2) were found to be in equilibrium in solution with the cyclic diamines HC[SiMe2N[(S)-1-phenylethyl]2](SiMe2NH-[(S)-1-phenylethyl]] (3) and HC[SiMe2NH[(R)-1-indanyl]][SiMe2NH[(R)-1-indanyl]) (4), which are generated upon ejection of one molecule of the chiral primary amine. Reaction of these equilibrium mixtures with three molar equivalents of butyllithium instantaneously gave the trilithium triamides HC[SiMe2N(Li)[(S)-1-phenylethyl]]3 (5) and HC[SiMe2N(Li)[(R)-1-indanyl]]3 (6), both of which were characterised by an X-ray diffraction study. Both lithium compounds possess a central heteroadamantane core, in which the two-coordinate Li atoms are additionally weakly solvated by the three aryl groups of the chiral peripheral substituents, the Li-C contacts being in the range of 2.65-2.73 A. Reaction of 5 and 6 with [TiCl4(thf)2] and ZrCl4 gave the corresponding amido complexes [TiCl-[HC[SiMe2N[(S)-1-phenylethyl]]3]] (7), [TiCl(HC[SiMe2N[(R)-1-indanyl]]3]] (8), [ZrCl[HC[SiMe2N[(S)-1-phenylethyl]]3]] (9) and [ZrCl[HC[SiMe2N[(R)-1-indanyl]]3]] (10), respectively. Of these, compound 7 was structurally characterised by X-ray structure analysis and was shown to possess a C3-symmetrical arrangement of the tripod ligand. The chiral anionic dinuclear complex [Li-(OEt2)4][Zr2Cl3[HC[SiMe2N[(S)-1-phenylethyl]]3]2] (11) was isolated from reaction mixtures leading to 9. An X-ray diffraction study established its dimeric structure, in which the chiral amido ligands cap the two metal centres, which are linked through three symmetrically arranged, bridging chloro ligands. Reaction of 9 and 10 with a series of alkyl Grignard and alkyllithium reagents yielded the corresponding alkylzirconium complexes. X-ray structure analyses of [Zr(CH3)[HC[SiMe2N[(S)-1-phenylethyl]]3]] (12) and [Zr(CH3)-[HC[SiMe2N)[(R)-1-indanyl]]3]] (20) established their detailed molecular arrangements. While the reaction of 12 with the aryl ketones PhC(O)R (R = CH = CHPh, iPr, Et) gave the corresponding C-O insertion products, which contain an additional chiral centre in the alkoxy group, with low stereoselectivity (0-40% de). The corresponding conversions with several aryl aldehydes yielded the alkoxo complexes with high stereoselectivity. Upon hydrolysis, the chiral alcohols were isolated and shown to have enantiomeric excesses between 68 and 82%. High stereodiscrimination was also observed in the insertion reactions of several chiral ketones and aldehydes. However, this was shown to originate primarily from the chirality of the substrate. In analogous experiments with carbonyl compounds, the ethyl- and butyl-zirconium analogues of 12 did not undergo CO insertion into the metal-alkyl bond. Instead, beta-elimination and formal insertion into the metal-hydride bond occurred. It was found that the elimination of the alkene was induced by  相似文献   

17.
《Electroanalysis》2003,15(22):1795-1798
Chiroselective binding sites have been created on thin gold films by application of the spreader‐bar approach. Impedometric techniques and surface plasmon resonance were applied to detect binding. (R)‐(+)‐1,1′‐Binaphthyl‐2,2′‐diol (R‐BNOH) and (S)‐(?)‐1,1′‐binaphthyl‐2,2′‐diol (S‐BNOH) were used as model analytes. The artificial receptors were prepared by co‐adsorption of 16‐mercaptohexadecane (matrix) with a thiol‐modified chiral selector (template). The conjugates of D ,L ‐thioctic acid and (R)‐(+)‐ or (S)‐(?)‐1,1′‐binaphthyl‐2,2′‐diamine were used as templates. Different concentration ratios of the matrix and template were tested. No chiral selectivity of surfaces formed by either the matrix or the template alone was observed. The use of alkylthiols shorter than 16‐mercaptohexadecane led to the formation of surfaces with no chiral selectivity. The gold electrodes coated by the spreader‐bar technique displayed an enantioselectivity of up to 4.76 or up to 2.55 as measured by the capacitive and SPR methods, respectively.  相似文献   

18.
Highly congested, axially chiral 1,8-bisphenolnaphthalenes have been synthesized in 75% overall yield by palladium-catalyzed Suzuki coupling of 1,8-diiodonaphthalene and 4-methoxy-2-methylphenylboronic acid followed by regioselective formylation and deprotection. The C(2)-symmetric anti-stereoisomers of 1,8-bis(2'-methyl-4'-hydroxy-5'-formylphenyl)naphthalene, 5, and its diimine analogues 9 and 10 were found to be significantly more stable than the corresponding syn-isomer. Crystallographic analysis revealed that this stereochemical preference results from a unique intramolecular hydrogen bonding motif and concomitant minimization of steric repulsion. Triaryl 5 proved stable to rotation about the chiral axes at room temperature and the enantiomers were isolated via formation of diastereomeric diimines with (R)-2-amino-1-propanol and (R)-2-amino-3-methyl-1-butanol, respectively, chromatographic separation, and mild hydrolysis. Slow syn/anti-interconversion of 5, 9, and 10 was observed at enhanced temperatures and the diastereomerization and enantiomerization processes were monitored by CD and NMR spectroscopy. The Gibbs activation energy, ΔG(?), for the isomerization of 5 was determined as 103.7 (102.4) kJ/mol for the conversion of the anti-(syn-) to the syn-(anti-)isomer at 45.0 °C. Condensation of 5 with two chiral amino alcohols generates diimines that undergo quantitative asymmetric transformation of the first kind toward the thermodynamically favored (P,P,R,R)- or (M,M,S,S)-atropisomer, respectively. The incorporation of two imino alcohol units controls the outcome of this unidirectional atropisomerization, i.e. the central chirality of the amino alcohol used induces a rigid, axially chiral triaryl scaffold with perfect stereocontrol. Accordingly, the rotational energy barrier for the conversion of (M,M,S,S)-9 to its syn-isomer is significantly increased and was determined as 115.7 kJ/mol at 58.0 °C.  相似文献   

19.
Kinetic resolution of acyclic secondary allylic silyl ethers by chiral dioxiranes generated in situ from chiral ketones (R)-1 and (R)-2 and Oxone was investigated. An efficient and catalytic method has been developed for kinetic resolution of those substrates with a CCl(3), tert-butyl, or CF(3) group at the alpha-position. In particular, high selectivities (S up to 100) were observed for kinetic resolutions of racemic alpha-trichloromethyl allylic silyl ethers 7 and 9-15 catalyzed by ketones (R)-2. Both the recovered substrates and the resulting epoxides were obtained in high enantiomeric excess. On the basis of steric and electrostatic interactions between the chiral dioxiranes and the racemic substrates, a model was proposed to rationalize the enantioselectivities and diastereoselectivities in the chiral ketone-catalyzed kinetic resolution process.  相似文献   

20.
High-resolution scanning tunneling microscopy has been used to examine the adsorbate structures formed when a racemic mixture of (9R,10R)-9,10-diiodooctadecan-1-ol and (9S,10S)-9,10-diiodooctadecan-1-ol is adsorbed at the basal plane of highly ordered pyrolytic graphite. The herringbone structure characteristic of the adsorption of long-chain molecules on graphite is observed. Close examination of the micrographs indicates a unique structure in which the chiral molecules adsorb in pairs, with one enantiomer filling half of the unit cell, and the other enantiomer filling the other half. Instead of forming separate chiral domains, as is sometimes observed when a racemic mixture adsorbs on an achiral surface, chiral pairs are formed and the pairs form an ordered monolayer, exposing opposite faces of the same molecule. An achiral racemic mixture is observed to form a chiral structure on an achiral surface in the regions of the surface examined here.  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