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 31 毫秒
1.
Novel [4, 6]helicenes ( 4a,b ) bearing a fused imidazolium unit have been prepared from [4, 6]helicene‐2,3‐di‐n‐propyl‐amines 3a,b . The in situ formation of N‐heterocyclic carbene (NHC) derivatives followed by their complexation to iridium(I) or rhodium(I) gave access to complexes 1a , 1′a , and 1b , containing mono‐coordinated helicene‐NHC, chloro and COD (COD = 1,5‐cyclooctadiene) ligands. Ir and Rh complexes 1a and 1′a were characterized by X‐ray crystallography. HPLC and NMR analyses showed that Ir(I) complex 1b existed as a mixture of two diastereomeric complexes corresponding to enantiomeric pairs M‐(?)/P‐(+)‐ 1b 1 and M‐(?)/P‐(+)‐ 1b 2 which differ by the position of COD through space. The chiroptical properties (electronic circular dichroism and optical rotation) of the four stereoisomers were measured. These complexes were also tested as catalysts in a transfer hydrogenation reaction.  相似文献   

2.
High‐performance liquid chromatographic methods were developed for the separation of the enantiomers of 19 β‐lactams. The direct separations were performed on chiral stationary phases containing either amylose‐tris‐3,5‐dimethylphenyl carbamate, (Kromasil® AmyCoat? column) or cellulose‐tris‐3,5‐dimethylphenyl carbamate, (Kromasil® CelluCoat? column) as chiral selector. The different methods were compared in systematic chromatographic examinations. The separations were carried out with good selectivity and resolution. The AmyCoat? and CelluCoat? columns appear to be highly complementary. The best separations of bi‐ and tricyclic β‐lactam stereoisomers were obtained with the AmyCoat? column, whereas the 4‐aryl‐substituted β‐lactams were better separated on the CelluCoat? column. The elution sequence was determined in all cases; no general rule could be established. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

3.
A simple one‐dimensional 13C NMR method is presented to discriminate between stereoisomers of organic compounds with more than one chiral center. By means of this method it is possible to discriminate between all eight stereoisomers of α‐tocopherol. To achieve this the chiral solvating agent (S)‐(+)‐1‐(9‐anthryl)‐2,2,2‐trifluoroethanol and the compound of interest were dissolved in high concentrations in chloroform‐d, and the nuclear magnetic resonance (NMR) spectrum was recorded at a low temperature. The individual stereoisomers of α‐tocopherol were assigned by spikes of the reference compounds. The method was also applied to six other representative examples. Chirality 27:850–855, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

4.
From achiral imidazolinium salts, chiral transition metal complexes containing an N-heterocyclic carbene (NHC) ligand were prepared (metal = palladium, copper, silver, gold, rhodium). Axial chirality in these complexes results from the formation of the metal-carbene bond leading to the restriction of rotation of dissymmetric N-aryl substituents about the C–N bond. When these complexes exhibited a sufficient configurational stability, a resolution by chiral high-performance liquid chromatography (HPLC) on preparative scale enabled isolation of enantiomers with excellent enantiopurities (>99% ee) and good yields. A study of the enantiomerization barriers revealed the effect of the backbone nature as well as the type of transition metal on its values. Nevertheless, the evaluation of palladium-based complexes in asymmetric intramolecular α-arylation of amides demonstrated that the ability to induce an enantioselectivity cannot be correlated to the configurational stability of the precatalysts.  相似文献   

5.
The origin of P‐ or M‐chirality of methyl substituted 1,3‐cyclohexadienes are elucidated by time‐dependent density functional theory (TD‐DFT) calculation of 1,3‐cyclohexadiene derivatives and acyclic 1,3‐dienes. The sign‐inversion of the rotatory strength of the lowest excited state between 1,3‐cyclohexadiene and (5R)‐axial‐methyl‐1,3‐cyclohexadiene is caused by the conformation around the (C=)C‐C(‐Me) dihedral angle. The correlation between the sign of the rotatory strength and conformation has been found not only in methyl substituted derivatives but also fluoro substituted compounds. Chirality 27:476–478, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
Bonded polysaccharide‐derived chiral stationary phases were found to be useful for the preparation of the four stereoisomers of the cyclopropane analogue of phenylalanine (c3Phe) as well as for the direct determination of the enantiomeric purity of c3Phe derivatives by HPLC. Three chiral stationary phases, consisting of cellulose and amylose derivatives chemically bonded on allylsilica gel, were tested. The mixed 10‐undecenoate/3,5‐dimethylphenylcarbamate of cellulose, 10‐undecenoate/3,5‐dimethylphenylcarbamate of amylose and 10‐undecenoate/p‐methylbenzoate of cellulose were the starting polysaccharide derivatives for CSP‐1, CSP‐2, and CSP‐3, respectively. Using mixtures of n‐hexane/chloroform/2‐propanol as mobile phase on a semi‐preparative column (150 mm × 20 mm ID) containing CSP‐2, we separated about 1.7 g of racemic cis‐methyl 1‐tert‐butoxycarbonylamino‐2‐phenylcyclopropanecarboxylate (cis‐ 6 ) and 1.2 g of racemic trans‐methyl‐1‐tert‐butoxycarbonylamino‐2‐phenylcycloprop‐anecarboxylate (trans‐ 6 ) by successive injections. Chirality 11:583–590, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

7.
The absolute configuration of 1 was deduced by vibrational circular dichroism together with the evaluation of the Flack and Hooft X‐ray parameters. Vibrational circular dichroism exciton coupling, using the carbonyl group signals, confirmed the absolute configuration of 2 . In addition, sodium borohydride reduction of the 11,13‐double bond of 6‐epi‐desacetyllaurenobiolide ( 1 ) yields an almost equimolecular mixture of C11 epimers, while reduction of the same double bond of 6‐epi‐laurenobiolide ( 2 ) provided almost exclusively the (11S) diastereoisomer 4 . Chirality 27:247–252, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

8.
Five ruthenium(II) complexes, i.e., [Ru(bpy)2(TIP)]2+ (bpy=2,2′‐bipyridine; TIP=2‐thiophenimidazo[4,5‐f] [1,10]phenanthroline; 1 ), [Ru(bpy)2(5‐NTIP)]2+ (5‐NTIP=2‐(5‐nitrothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 2 ), [Ru(bpy)2(5‐MOTIP)]2+ (5‐MOTIP=2‐(5‐methoxythiophen)imidazo[4,5‐f] [1,10]phenanthroline; 3 ), [Ru(bpy)2(5‐BTIP)]2+ (5‐BTIP=2‐(5‐bromothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 4 ), and [Ru(bpy)2(4‐BTIP)]2+ (4‐BTIP=2‐(4‐bromothiophen)imidazo[4,5‐f] [1,10]phenanthroline; 5 ), were synthesized and characterized by elemental analysis and UV/VIS, IR, and 1H‐NMR spectroscopic methods. The photophysical and DNA‐binding properties were investigated by means of UV and fluorescence spectroscopic methods and viscosity measurements, respectively. The results suggest that all five complexes can bind to CT‐DNA with various binding strength. Complexes 2 and 3 showed the strongest and the weakest binding affinity, respectively, among these five complexes. Due to the substituent position of the Br‐atom in the ligand, complex 5 interacted stronger with CT‐DNA than complex 4 . The binding affinities of the complexes decreased in the order 2, 5, 4, 1 , and 3 .  相似文献   

9.
Tetrapodal ligands H4L1 and H4L2 containing imidazole groups have been synthesized by the reaction of 1,10‐phenanthroline‐5,6‐dione with 1,2,4,5‐tetrakis[(4‐formylphenoxy)methyl]benzene and 1,2,4,5‐tetrakis[(3‐formylphenoxy)methyl]benzene, respectively, in presence of NH4OAc. Two star‐shaped complexes [{Ru(bpy)2}44‐H4L1)](PF6)8 and [{Ru(bpy)2}44‐H4L2)](PF6)8 (bpy = 2,2′‐bipyridine) have been prepared by refluxing Ru(bpy)2Cl2·2H2O and each ligand in ethylene glycol. The deprotonated complexes [{Ru(bpy)2}44‐L1)](PF6)4 and [{Ru(bpy)2}44‐L2)](PF6)4 have been obtained by the reaction of sodium methoxide with [{Ru(bpy)2}44‐H4L1)](PF6)8 and [{Ru(bpy)2}44‐H4L2)](PF6)8, respectively, in methanol. The pH effects on the UV–vis light absorption and emission spectra of both complexes have been studied, and ground‐ and excited‐state ionization constants of both complexes have been derived. The photophysical properties of both complexes are strongly dependent on the solution pH. They act as proton‐induced off–on–off luminescent sensors through two successive deprotonation processes of imidazole groups, with a maximum on–off ratio of 8 in buffer solution at room temperature. Theoretical calculations for the highest occupied molecular orbital (HOMO) and lowest occupied molecular orbital (LOMO) orbitals of bridging ligand are also presented for plausible explanations of the fluorescence changes. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

10.
Reproductive barriers play an important role in the maintenance of species boundaries. However, to date, few studies have provided a detailed analysis of reproductive isolation barriers between species or examined their importance in maintaining species identity. This is the first detailed study into pre‐ and post‐zygotic reproductive isolation barriers in Antirrhinum, based on a mixed population with two species that rarely co‐occur. The study revealed that pollinator constancy and preference and poor hybrid seed viability were the most important reproductive isolating mechanisms. Reproductive isolation was practically complete by both pre‐ and post‐zygotic barriers. Average pre‐zygotic isolation was greater than post‐zygotic isolation, in accordance with the trend observed in flowering plants in which reproductive isolation is principally caused by pre‐zygotic mechanisms. However, average post‐zygotic isolation was also high, in contrast to what was expected among Antirrhinum spp. This case highlights the importance of quantifying the reproductive isolation barriers thoroughly to understand how and why species boundaries are maintained. © 2014 The Linnean Society of London, Botanical Journal of the Linnean Society, 2014, 176 , 159–172.  相似文献   

11.
The new complex compounds [RuLCl(p‐cymene)] ? 3H2O and [NiL2(H2O)2] ? 3H2O (L: 1‐{4‐[(2‐hydroxy‐3‐methoxybenzylidene)amino]phenyl}ethanone) were prepared and characterized using FT‐IR, 1H‐ and 13C‐NMR, mass spectroscopy, TGA, elemental analysis, X‐ray powder diffraction and magnetic moment techniques. Octahedral geometry for new Ni(II) and Ru(II) complexes was proposed. Thermal decomposition confirmed the existence of lattice and coordinated water molecule in the complexes. To determine the antioxidant properties of Schiff base ligand and its Ni(II), Ru(II) metal complexes, FRAP, CUPRAC, ABTS and DPPH methods of antioxidant assays were used. Moreover, enzyme inhibition of complexes was evaluated against carbonic anhydrase I and II isoenzymes (CA I and CA II) and acetylcholinesterase (AChE). For CA I and CA II, the best inhibition enzymes, was the Ni(II) complex with 62.98±18.41, 86.17±23.62 Ki values, whereas this inhibition effect showed ligand with 24.53±2.66 Ki value for the AChE enzyme.  相似文献   

12.
Female European corn borer, Ostrinia nubilalis, produce and males respond to sex pheromone blends with either E‐ or Z‐Δ11‐tetradecenyl acetate as the major component. E‐ and Z‐race populations are sympatric in the Eastern United States, Southeastern Canada, and the Mediterranean region of Europe. The E‐ and Z‐pheromone races of O. nubilalis are models for incipient species formation, but hybridization frequencies within natural populations remain obscure due to lack of a high‐throughput phenotyping method. Lassance et al. previously identified a pheromone gland‐expressed fatty‐acyl reductase gene (pgfar) that controls the ratio of Δ11‐tetradecenyl acetate stereoisomers. We identified three single nucleotide polymorphism (SNP) markers within pgfar that are differentially fixed between E‐ and Z‐race females, and that are ≥98.2% correlated with female pheromone ratios measured by gas chromatography. Genotypic data from locations in the United States demonstrated that pgfar‐z alleles were fixed within historically allopatric Z‐pheromone race populations in the Midwest, and that hybrid frequency ranged from 0.00 to 0.42 within 11 sympatric sites where the two races co‐occur in the Eastern United States (mean hybridization frequency or heterozygosity (HO) = 0.226 ± 0.279). Estimates of hybridization between the E‐ and Z‐races are important for understanding the dynamics involved in maintaining race integrity, and are consistent with previous estimates of low levels of genetic divergence between E‐ and Z‐races and the presence of weak prezygotic mating barriers.  相似文献   

13.
Guozhen Wu  Peijie Wang 《Chirality》2015,27(11):820-825
A bond polarizability algorithm was developed and applied to interpret the Raman optical activity (ROA) intensity. It is demonstrated that for the chiral molecule such as S(+)2,2‐dimethyl‐1,3‐dioxolane‐4‐methanol there exists approximate (or symmetry breaking) mirror reflection that reverses the signs of the differential bond polarizabilities of the pair bond coordinates that are related to each other by the mirror reflection, just like that between the right and left enantiomers. The magnitude difference of the differential bond polarizabilities of the pair bond coordinates becomes smaller as they are farther away from the asymmetric atom. Hence, that the asymmetric atom (center) plays a central role in ROA is confirmed from a spectroscopic viewpoint. Meanwhile, the concept of intramolecular enantiomerism is proposed. Chirality 27:820–825, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

14.
A reaction of DBU promoted ring opening in nucleoside‐3'‐O‐ and nucleoside‐5'‐O‐(2‐thio‐4,4‐pentamethylene‐1,3,2‐oxathiaphospholane) monomers with a pyrophosphate or a methylenediphosphonate anion proceeds with substantial loss of stereoselectivity. Depending on the absolute configuration of the phosphorus atom, so far widely accepted the stereoretentive mechanism of condensation is accompanied by a stereoinvertive one, most likely employing an intramolecular ligand–ligand exchange in an uncharged intermediate. Chirality 27:155–122, 2015. © 2014 Wiley Periodicals, Inc  相似文献   

15.
The four stereoisomers of the antimuscarinic 3-(2,3-dihydrobenzofuran-2-yl)quinuclidine have been prepared by a method involving chromatographic separation of the racemic diastereoisomers as borane complexes. The relative and absolute configurations of the stereoisomers were determined by X-ray crystallographic methods. The crystal structure of (2′R,3R)-3-(2,3-dihydrobenzofuran-2-yl)quinuclidine · HCl · H2O contains two independent molecules with different conformations of both the quinuclidine moiety and the dihydrofuran ring. Chirality 10:813–820, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

16.
Hydroxynitrile lyases (HNLs) catalyze the conversion of chiral cyanohydrins to hydrocyanic acid (HCN) and aldehyde or ketone. Hydroxynitrile lyase from Arabidopsis thaliana (AtHNL) is the first R‐selective HNL enzyme containing an α/β‐hydrolases fold. In this article, the catalytic mechanism of AtHNL was theoretically studied by using QM/MM approach based on the recently obtained crystal structure in 2012. Two computational models were constructed, and two possible reaction pathways were considered. In Path A, the calculation results indicate that the proton transfer from the hydroxyl group of cyanohydrin occurs firstly, and then the cleavage of C1‐C2 bond and the rotation of the generated cyanide ion (CN?) follow, afterwards, CN? abstracts a proton from His236 via Ser81. The C1‐C2 bond cleavage and the protonation of CN? correspond to comparable free energy barriers (12.1 vs. 12.2 kcal mol?1), suggesting that both of the two processes contribute a lot to rate‐limiting. In Path B, the deprotonation of the hydroxyl group of cyanohydrin and the cleavage of C1‐C2 bond take place in a concerted manner, which corresponds to the highest free energy barrier of 13.2 kcal mol?1. The free energy barriers of Path A and B are very similar and basically agree well with the experimental value of HbHNL, a similar enzyme of AtHNL. Therefore, both of the two pathways are possible. In the reaction, the catalytic triad (His236, Ser81, and Asp208) acts as the general acid/base, and the generated CN? is stabilized by the hydroxyl group of Ser81 and the main‐chain NH‐groups of Ala13 and Phe82. Proteins 2015; 83:66–77. © 2014 Wiley Periodicals, Inc.  相似文献   

17.
The preparation of all four stereoisomers of the proline analog that bears a phenyl group attached to the β carbon either cis or trans to the carboxylic acid (cis‐ and trans‐β‐phenylproline, respectively) has been addressed. The methodology developed allows access to multigram quantities of the target amino acids in enantiomerically pure form and suitably protected for use in peptide synthesis. Racemic precursors of cis‐β‐phenylproline and trans‐β‐phenylproline were prepared from easily available starting materials and subjected to high‐performance liquid chromatography enantioseparation. Semipreparative columns (250 × 20 mm) containing chiral stationary phases based on amylose (Chiralpak IA) (Daicel‐Chiral Technologies Europe, Illkirch, France) or cellulose (Chiralpak IC) were used respectively for the resolution of the cis‐ and trans‐β‐phenylproline precursors. Chirality, 24:1082‐1091, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

18.
Permethrin (PM), cypermethrin (CP), and cyfluthrin (CF) are three important synthetic pyrethroids, which contain two, four, and four enantiomeric pairs (diastereomers) and thus have four, eight, and eight stereoisomers, respectively. In this study, the stereo‐ and enantioselective degradation of PM, CP, and CF in a Shijiazhuang alkaline yellow soil and a Wuhan acidic red soil were studied in detail by a combination of achiral and chiral high‐performance liquid chromatography (HPLC). The results showed that PM, CP, and CF degraded faster in Shijiazhuang soil than in Wuhan soil, and the dissipation rate followed an order of PM > CF > CP in both soils. The three pyrethroids exhibited similar diastereomer selectivity, while CP and CF showed higher enantioselectivity than PM. Moreover, the trans‐diastereomers degraded faster, and showed higher enantioselectivity than the corresponding cis‐diastereomers. For PM, the enantiomer 1S‐trans‐PM degraded most rapidly in both soils. As for CP and CF, the highest enantioselectivity was observed for diastereomer trans‐3, and the insecticidally active enantiomer 1R‐trans‐αS degraded fastest among the 8 CP or CF stereoisomers in both soils. In addition, the Wuhan acidic soil displayed higher diastereomer and enantiomer selectivity than the Shijiazhuang alkaline soil for the three pyrethroids. Further incubation of CF in an alkaline‐treated Wuhan soil showed that the dissipation rate greatly increased and the diastereo‐ and enantioselectivity significantly decreased after the alkaline treatment process. Chirality 28:72–77, 2016. © 2015 Wiley Periodicals, Inc.  相似文献   

19.
An important evolutionary question concerns whether one or many barriers are involved in the early stages of speciation. We examine pre‐ and post‐zygotic reproductive barriers between two species of butterflies (Heliconius erato chestertonii and H. e. venus) separated by a bimodal hybrid zone in the Cauca Valley, Colombia. We show that there is both strong pre‐ and post‐mating reproductive isolation, together leading to a 98% reduction in gene flow between the species. Pre‐mating isolation plays a primary role, contributing strongly to this isolation (87%), similar to previous examples in Heliconius. Post‐mating isolation was also strong, with absence of Haldane’s rule, but an asymmetric reduction in fertility (< 11%) in inter‐specific crosses depending on maternal genotype. In summary, this is one of the first examples of post‐zygotic reproductive isolation playing a significant role in early stages of parapatric speciation in Heliconius and demonstrates the importance of multiple barriers to gene flow in the speciation process.  相似文献   

20.
Both enantiomers of three biologically relevant paraconic acids—MB‐3, methylenolactocin, and C75—were obtained with enantioselectivities up to 99% by kinetic enzymatic resolutions. Good enantiomeric excesses were obtained for MB‐3 and methylenolactocin, using α‐chymotrypsin and aminoacylase as enantiocomplementary enzymes, while C75 was resolved with aminoacylase. They all were evaluated for their antiproliferative, antibacterial, and antifungal activities, showing weak effects and practically no difference between enantiomers in each case. At high concentrations (16–64 µg/mL), (–)‐ C75 acted as an antimicrobial agent against Gram‐positive bacteria. Chirality 27:239–246, 2015. © 2015 Wiley Periodicals, Inc.  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