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.
A possible role that might have been played by ordered clusters at the air/water interface for the generation of homochiral oligopeptides under prebiotic conditions has been probed by a catalyzed polymerization of amphiphilic activated alpha-amino acids that assembled as two-dimensional (2-D) crystallites at this interface. Three type of processes are described: (i) polymerization of racemates of activated alpha-amino acids that undergo spontaneous resolution into enantiomorphous 2-D crystallites to yield racemic mixtures of oligopeptides enriched with the oligomers of homochiral sequence, (ii) enhanced formation of racemic mixtures of homochiral oligopeptides via lattice-controlled polymerization within 2-D racemic compounds and (iii) generation of homochiral oligopeptides of a single handedness from chiral non-racemic mixtures of monomers that self-assemble into two different phases, racemic crystallites composed from both enantiomers and enantiomorphous crystallites of the enantiomer in excess. The structures of the 2-D crystallites have been determined by grazing incidence X-ray diffraction and the diastereoisomeric composition of the oligopeptides by matrix-assisted laser-desorption time-of-flight mass spectrometry with enantio-labeling.  相似文献   

2.
Thioesters of α-amino acids are considered as plausible monomers for the generation of the primeval peptides. DL-Leucine-thioethyl esters (LeuSEt), where the L-enantiomer was tagged with deuterium atoms, undergo polycondensation in water or in bicarbonate or imidazole buffer solutions to yield mainly heterochiral (atactic) peptides and diketopiperazine, as analyzed by MALDI-TOF and ESI mass-spectrometry. In variance, when polymerization of DL(d10)-Leu, first activated with N,N′-carbonyldiimidazole, then initiated with ethanethiol or with DL(d3)-LeuSEt yielded a library of peptides up to 30 detectable residues where those of homochiral sequence (isotactic) are the dominant diastereoisomers. At these conditions, racemic β-sheets are formed and operate as stereoselective templates in the process of chain-elongation. Isotopic L:L(d10)-Leu co-peptides were obtained in the polymerization of L(d10)-Leu with L-LeuSEt. By contrast, mixtures of oligo-D-Leu and oligo-L(d10)-Leu were obtained in the polymerization of mixtures of D-LeuSEt with activated L(d10)-Leu. Isotactic co-peptides containing Leu and Val residues were formed in the polymerization of mixtures of activated DL(d8)-Val with DL(d3)-LeuSEt in water, implying that the racemic β-sheets exert regio-enantio-selection but not chemo-selection. A reaction pathway is suggested, where LeuSEt operates both as initiator of the reaction as well as a multimer.  相似文献   

3.
Experimental results show that benzil (1,2‐diphenyl‐1,2‐ethanedione), an achiral compound that crystallizes as a racemic conglomerate, yields by solidification polycrystalline scalemic mixtures of high enantiomeric excesses. These results are related to those previously reported in this type of compounds on deracemizations of racemic mixtures of crystal enantiomorphs obtained by wet grinding. However, the present results strongly suggest that these experiments cannot be explained without taking into account chiral recognition interactions at the level of precritical clusters. The conditions that would define a general thermodynamic scenario for such deracemizations are discussed. Chirality 25:393–399, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
The resolution of racemic ibuprofen was studied by partial diastereomer salt formation. The resolution was performed via two methods: resolution with (+)-(R)-phenylethylamine as chiral agent and resolution with a mixture of (+)-(R)-phenylethylamine and benzylamine. The diastereomers and unreacted enantiomers were separated by supercritical fluid extraction with carbon dioxide at 15 MPa and 33 degrees C. The influence of the achiral benzylamine on the resolution efficiency was studied by varying the concentrations of the structurally related amines in their mixtures, keeping the sum molar ratio of the amines to racemic ibuprofen constant at 0.55 +/- 0.02. The presence of benzylamine positively influenced the resolution efficiency at certain concentrations. The crystal structure of the salts of (+)-(R)-phenylethylamine with (-)-(R)-ibuprofen and (+)-(S)-ibuprofen, respectively, as well as the cocrystal of the benzylamine-ibuprofen salt with neutral ibuprofen molecules are presented. These structures were determined by single crystal X-ray diffraction, proving the significantly different stoichiometry of the related amines with the chiral acid, in accordance with mass balance calculations.  相似文献   

5.
Crystalline characteristics of racemic, pure R and S enantiomers and physical mixtures of Ketoprofen (KET) have been studied by DSC and X‐ray diffractometry. Aqueous solubilities were 182.6 ± 9.1 μg/ml for racemic KET, 259.6 ± 6.6 μg/ml for R‐KET, and 304.3 ± 2.7 μg/ml for S‐KET. Matrix tablets made with racemic and physical mixtures of KET show stereoselective drug release, which is faster for S‐KET than for R‐KET. This effect is more marked when the chiral excipient hydroxypropylmethylcellulose (HPMC) is used in place of the achiral Eudragit RL. Stereoselectivity of release is also affected by the amount of KET. Similar results were obtained when another chiral drug with low solubility, Ricobendazole (RBZ), is used. Depending on the excipient and drug dosage, more or less marked stereoselective drug release is obtained in RBZ matrix tablet formulations. Chirality 11:611–615, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

6.
This article is concerned with the spontaneous onset of homochiral oligopeptide sequences. We will show that the polymerization of hydrophobic NCA (N-carboxyanhydride = cyclic anhydride)-amino acid racemates (i.e. tryptophane, leucine and isoleucine) in aqueous solution yields oligopeptides that are characterized by a high degree of homochiral sequences. Furthermore we will show that quartz enhances efficiently the mole fraction of oligopeptides with homochiral sequence by selectively adsorbing the more stereoregular oligopeptides from an aqueous solution of oligo-D,L-leucine. We find in particular that the mole fraction of the adsorbed homochiral 7mers is 17 times larger than the mole fraction calculated for a theoretical, random process. Experimentally the stereoisomer distribution for each oligomer length can be determined by the use of enantio-labeling and LC-MS (Liquid Chromatography-Mass Spectrometry). Furthermore, if we start the polymerization with an enantiomeric excess (e.e.) of 20% of L-leucine (L-amino acid: D-amino acid = 6:4, molar ratio) we observe a chiral amplification in the enantiomeric homochiral oligopeptides. We think that such processes are relevant to the chemical evolution of single handedness.  相似文献   

7.
Amino acid auxotrophous bacteria such as Lactococcus lactis use proteins as a source of amino acids. For this process, they possess a complex proteolytic system to degrade the protein(s) and to transport the degradation products into the cell. We have been able to dissect the various steps of the pathway by deleting one or more genes encoding key enzymes/components of the system and using mass spectrometry to analyse the complex peptide mixtures. This approach revealed in detail how L . lactis liberates the required amino acids from β-casein, the major component of the lactococcal diet. Mutants containing the extracellular proteinase PrtP, but lacking the oligopeptide transport system Opp and the autolysin AcmA, were used to determine the proteinase specificity in vivo . To identify the substrates of Opp present in the casein hydrolysate, the PrtP-generated peptide pool was offered to mutants lacking the proteinase, but containing Opp, and the disappearance of peptides from the medium as well as the intracellular accumulation of amino acids and peptides was monitored in peptidase-proficient and fivefold peptidase-deficient genetic backgrounds. The results are unambiguous and firmly establish that (i) the carboxyl-terminal end of β-casein is degraded preferentially despite the broad specificity of the proteinase; (ii) peptides smaller than five residues are not formed in vivo  ; (iii) use of oligopeptides of 5–10 residues becomes only possible after uptake via Opp; (iv) only a few (10–14) of the peptides generated by PrtP are actually used, even though the system facilitates the transport of oligopeptides up to at least 10 residues. The technology described here allows us to monitor the fate of individual peptides in complex mixtures and is applicable to other proteolytic systems.  相似文献   

8.
PH. Dumas  P. Sigwalt 《Chirality》1991,3(6):484-491
The polymerization of racemic methylthiirane in homogeneous phase, initiated by bis(isopropyl-S-cysteinato) cadmium is a living process. The resulting polymers are isotactic and optically active at partial conversion. The optical purity of the residual monomer may reach 27% at half conversion. The propagation occurs mainly on one valency of Cd, however oligomers grow slowly on the second valency. The stereoregularity of the polymer chain appears only when the length of the oligomer becomes high enough, making possible a bicoordination of the Cd counterions. The stereoregularity of the polymer is characterized by the molar fraction σ of isotactic diads which varies from 0.5 for atactic chains—formed at the beginning—to about one for isotactic segments formed for longer chains. The stereospecifictity also depends on temperature of propagation and on initiator concentration. The kinetics observed (zero order in monomer and one-half in Cd) are explained by monomer coordination before insertion and dimeric association of the thiolate end groups. The enantioasymmetric process observed results from an unbalance in the number of the two different types of active sites and possibly from a difference in their reactivities. Enantioasymmetry has been found to decrease significantly when the dielectric constant ε of the medium increases.  相似文献   

9.
β-methylaspartate ammonia-lyase, EC 4.3.1.2, (β-methylaspartase) from Clostridium tetanomorphum was used to produce a 40/60 molar ratio of (2S,3R) and (2S,3S)-3-methylaspartic acids, 2a and 2b , respectively, from mesaconic acid 1 as substrate, on a large scale. To prepare (3R,4R)-3-methyl-4-(benzyloxycarbonyl)-2-oxetanone (benzyl 3-methylmalolactonate) 6, 2a and 2b were transformed, in the first step, into 2-bromo-3-methylsuccinic acids 3a and 3b and separated. After three further steps, (2S,3S)- 3a yielded the α,β-substituted β-lactone (3R,4R) 6 with a very high diastereoisomeric excess (>95% by chiral gas chromatography). The corresponding crystalline polymer, poly[benzyl β-(2R,3S)-3-methylmalate] 8 , prepared by an anionic ring opening polymerization, was highly isotactic as determined by 13C NMR. Catalytic hydrogenolysis of lactone 6 yielded (3R,4R)-3-methyl-4-carboxy-2-oxetanone (3-methylmalolactonic acid) 7 , to which reactive, chiral, or bioactive molecules can be attached through ester bonds leading to polymers with possible therapeutic applications. Because of the ability of β-methylaspartase to catalyse both syn- and anti-elimination of ammonia from (2S,3RS)-3-methylaspartic acid 2ab at different rates, the (2S,3R)-stereoisomer 2a was retained and isolated for further reactions. These results permit the use of the chemoenzymatic route for the preparation of both optically active and racemic polymers of 3-methylmalic acid with well-defined enantiomeric and diastereoisomeric compositions. Chirality 10:727–733, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
Zhang P  Polavarapu PL  Huang J  Li T 《Chirality》2007,19(2):99-105
A chiral column, with decaproline as the chiral selector, has broad chiral selectivity. To understand the separation mechanism of this chiral column, multiple spectroscopic techniques, including optical rotation, electronic circular dichroism, infrared absorption and vibrational circular dichroism, have been used here to study the conformation of the decaproline oligomer in isopropanol(IPA)/dichloromethane(DCM) mixtures. These studies indicate that decaproline oligomer adopts polyproline II conformation in IPA/DCM solvent system (0% IPA approximately 100% IPA). Hydrogen bonding interactions between C=O groups of decaproline and IPA molecules increase as the content of IPA in the solvent mixture increases up to 60% and become less significant from then onwards. These spectroscopic observations are found to have a good correlation with the enantiomeric separation of racemic 2,2,2-trifluoro-1-[10-(2,2,2-trifluoro-1-hydroxy-ethyl-anthracen-9-yl]-ethanol by the decaproline column.  相似文献   

11.
Kuroda R  Imai Y  Sato T 《Chirality》2001,13(9):588-594
New adduct crystals were obtained by simply mixing/grinding component crystals of bis-beta-naphthol (BN) derivatives with benzoquinone (BQ) under solvent-free conditions. Chiral recognition was found to operate during this process and either a racemic or a chiral crystal of a BN derivative produced an adduct crystal with BQ by solid-state crystallization. The chirality preference changed subtly according to the molecular structure of the BN derivative. Even in circumstances in which no adduct was formed, addition of a third component, such as crystals of naphthalene, to the grinding mixture yielded an adduct crystal. Remarkably, these adduct crystals were found to decompose spontaneously with time and revert to the starting crystal of the BN derivative by losing BQ molecules from the crystal lattice. Local melting of crystals by the grinding pressure was found unlikely to be the mechanism of adduct formation. Overall, these results demonstrate that molecules in the solid state could change their relative location and hydrogen bonding partners, thereby exerting chiral discrimination.  相似文献   

12.
Usami Y  Okada Y  Yamada T 《Chirality》2011,23(Z1):E7-11
Pericosines are carbasugar-type metabolites of Periconia byssoides OUPS-N133, a fungus that was originally separated from the sea hare Aplysia kurodai. It has been reported that pericosines C and E are enantiomeric mixtures. The difference in specific rotation between natural pericosine C and the synthetic one led to the conclusion that natural pericosine C is an enantiomeric mixture. Meanwhile, the small specific rotation of natural pericosine B compared to that of the synthetic one led to the deduction that natural pericosine B might also be an enantiomeric mixture. Then, racemic pericosines B and C were synthesized, and the direct enantioseparation of these racemic carbasugars was conducted with CHIRALPAK? IA and CHIRALPAK? AY-H, which are suitable columns for racemic pericosines B and C, respectively. Using chiral HPLC, we conclude that natural pericosines B and C exist as enantiomeric mixtures. A rare example of the application of direct chiral HPLC analysis to intact sugars or carbasugars was provided.  相似文献   

13.
Structure–interaction relationships, stereoselectivity, and solubility enhancement in inclusion compexation of β-cyclodextrins (CDs) with some racemic and enantiomerically pure 1,4-dihydropyridine derivatives (DHPs) were investigated. 1:1 and 1:2 (mole ratio) complexes were prepared and characterized by X-ray powder diffraction, differential scanning calorimetry (DSC), MS-FAB spectrometry, 1H-NMR spectroscopy, water and phase solubility. The solubility studies have revealed different complexation equilibria for optically pure DHP enantiomers, and corresponding racemic mixtures in water solutions. By means of 1H-NMR chemical shift measurements, the inclusion of aromatic fragments of racemic and enantiomerically pure DHP molecules within the cavities of different CDs was elucidated. Considerable stereoselectivity in complexation interactions was observed. The results indicate the potential use of cyclodextrins as chiral selectors for enantiomeric resolution of 1,4-DHP calcium antagonists. © 1993 Wiley-Liss, Inc.  相似文献   

14.
Aminopiperazinone inhibitors of BACE were identified by rational design. Structure based design guided idea prioritization and initial racemic hit 18a showed good activity. Modification in decoration and chiral separation resulted in the 40 nM inhibitor, (−)-37, which showed in vivo reduction of amyloid beta peptides. The crystal structure of 18a showed a binding mode driven by interaction with the catalytic aspartate dyad and distribution of the biaryl amide decoration towards S1 and S3 pockets.  相似文献   

15.
A fundamental understanding of the enantiospecific interactions between chiral adsorbates and understanding of their interactions with chiral surfaces is key to unlocking the origins of enantiospecific surface chemistry. Herein, the adsorption and decomposition of the amino acid proline (Pro) have been studied on the achiral Cu(110) and Cu(111) surfaces and on the chiral Cu(643)R&S surfaces. Isotopically labelled 1-13C-l- Pro has been used to probe the Pro decomposition mechanism and to allow mass spectrometric discrimination of d -Pro and 1-13C-l -Pro when adsorbed as mixtures. On the Cu(111) surface, X-ray photoelectron spectroscopy reveals that Pro adsorbs as an anionic species in the monolayer. On the chiral Cu(643)R&S surface, adsorbed Pro enantiomers decompose with non-enantiospecific kinetics. However, the decomposition kinetics were found to be different on the terraces versus the kinked steps. Exposure of the chiral Cu(643)R&S surfaces to a racemic gas phase mixture of d -Pro and 1-13C-l -Pro resulted in the adsorption of a racemic mixture; i.e., adsorption is not enantiospecific. However, exposure to non-racemic mixtures of d -Pro and 1-13C-l -Pro resulted in amplification of enantiomeric excess on the surface, indicative of homochiral aggregation of adsorbed Pro. During co-adsorption, this amplification is observed even at very low coverages, quite distinct from the behavior of other amino acids, which begin to exhibit homochiral aggregation only after reaching monolayer coverages. The equilibrium adsorption of d -Pro and 1-13C-l -Pro mixtures on achiral Cu(110) did not display any aggregation, consistent with prior scanning tunneling microscopy (STM) observations of dl -Pro/Cu(110). This demonstrates convergence between findings from equilibrium adsorption methods and STM experiments and corroborates formation of a 2D random solid solution.  相似文献   

16.
The application of cellulose-based stationary phases for chiral separations has been extended to open tubular column chromatography. Efficient columns were obtained by coating the capillaries with mixtures of chiral cellulose materials and conventional achiral stationary phases for gas chromatography. In this study, various siloxane and polyethylene glycol polymers were used as achiral components and mixed with different substituted benzoylcellulose derivatives as chiral components. Systematic investigations were carried out to determine the optimal ratio for the components of the stationary phase. Depending on the chromatographic mode—gas chromatography (GC) or supercritical fluid chromatography (SFC)—the stationary phases were found to behave differently. The applicability of the technique was demonstrated by the resolution of various racemic compounds. © 1993 Wiley-Liss, Inc.  相似文献   

17.
Moussa A  Pham C  Bommireddy S  Muller G 《Chirality》2009,21(5):497-506
The perturbation of the racemic equilibrium of luminescent D3 terbium(III) complexes with chelidamic acid (CDA), a hydroxylated derivative of 2,6-pyridine-dicarboxylic acid (DPA), by added chiral biomolecules such as L-amino acids has been studied using circularly polarized luminescence and 13C NMR spectroscopy. It is shown in this work that the chiral-induced equilibrium shift of [Tb(CDA)3](6-) by L-amino acids (i.e. L-proline or L-arginine) was largely influenced by the hydrogen-bonding networks formed between the ligand interface of racemic [Tb(CDA)3](6-) and these added chiral agents. The capping of potential hydrogen-bonding sites by acetylation in L-proline led to a approximately 100-fold drop in the induced optical activity of the [Tb(CDA)3](6-):N-acetyl-L-proline system. This result suggested that the hydrogen-bonding networks serve as the basis for further noncovalent discriminatory interactions between racemic [Tb(CDA)3](6-) and added L-amino acids.  相似文献   

18.
Jiao F  Yang W  Wang F  Tian L  Li L  Chen X  Mu K 《Chirality》2012,24(8):661-667
A method of solvent sublation was developed for the enantioseparation of racemic ofloxacin (rac Oflx) and racemic tryptophan (rac Trp). In this method, dibenzoyl-L-tartaric acid (L-DBTA) and di-(2-ethylhexyl) phosphoric acid (D2EHPA) and sodium lauryl sulfate (SDS) were used as chiral coextractants and foamer, respectively. Several important parameters influencing the separation performances, such as pH in aqueous phase, concentrations of rac mixtures, L-DBTA, D2EHPA, and SDS, were investigated. Under the optimal operation conditions, the enantiomeric excess and enantioselectivity were 60.08% and 5.58 for Oflx and 65.09% and 6.31 for Trp, respectively. The yields of D-enantiomer and L-enantiomer were 34.23% and 8.54% for Oflx and 18.59% and 3.93% for Trp, respectively. The results suggest that the enantioselectivities have been enhanced compared with the traditional chiral extraction. This technique is an efficient chiral separation method, with many advantages such as low expenditures of organic solvent, low consumption of chiral extractant, and easy realization of multistage operation.  相似文献   

19.
The binary phase diagrams of hydrogen halides salts of medetomidine (Med.HX, X:Br,I) and hydrogen oxalate salt of medetomidine (Med.Ox) were determined based on thermogravimetric/differential thermal analysis (TGA/DTA) and their crystal structure behavior was confirmed by comparison of the X‐ray diffractometry and FT‐IR spectroscopy of the racemate and pure enantiomer. All hydrogen halide salts presented racemic compound behavior. Heat of fusion of halides salt of (rac)‐medetomidine decreased with ionic radius increase. Eutectic points for Med.HCl (previously reported), Med.HBr, and Med.HI rest were unchanged approximately. The solubility of different enantiomeric mixtures of Med.HBr and Med.HI were measured at 10, 20, and 30°C in 2‐propanol showing a solubility increase with ionic radius. A binary phase diagram of Med.Ox shows a racemic conglomerate behavior. The solubility of enantiomeric mixtures of Med.Ox were measured at 10, 20, 30, and 40°C. The ternary phase diagram of Med.Ox in ethanol conforms to a conglomerate crystal forming system, favoring its enantiomeric purification by preferential crystallization. Chirality 26:183–188, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
The organic compounds synthesized in prebiotic experiments are racemic mixtures. A number of proposals have been offered to explain how asymmetric organic compounds formed on the Earth before life arose, with the influence of chiral weak nuclear interactions being the most frequent proposal. This and other proposed asymmetric syntheses give only slight enantiomeric excess and any slight excess will be degraded by racemization. This applies particularly to amino acids where half-lives of 10(5)-10(6) years are to be expected at temperatures characteristic of the Earth's surface. Since the generation of chiral molecules could not have been a significant process under geological conditions, the origins of this asymmetry must have occurred at the time of the origin of life or shortly thereafter. It is possible that the compounds in the first living organisms were prochiral rather than chiral; this is unlikely for amino acids, but it is possible for the monomers of RNA-like molecules.  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