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 125 毫秒
1.
Pasteur carried out pioneering work on cinchona alkaloids and their derivatives and his studies led to important discoveries. He examined crystals of cinchona alkaloids for his correlation of crystal hemihedrism with molecular chirality, studies that led Pasteur to the discovery of physicochemical differences between diastereoisomeric salts of tartaric acids with optically active cinchona bases, an important insight into fundamentals of molecular chirality. These physicochemical differences also led to Pasteur’s invention of the vital method of racemate resolution through diastereoisomeric derivatives. Pasteur clarified the confusion around the cinchona alkaloids by elucidating their identities and relations. He discovered the conversion of the major cinchona alkaloids to quinicine and cinchonicine, a finding subsequently of considerable importance in studies of the structure and synthesis of the major cinchona alkaloids. The reaction producing quinicine and cinchonicine led Pasteur to the discovery of the racemization of tartaric acid and the finding of meso‐tartaric acid, fundamental breakthroughs in the development of stereochemistry.  相似文献   

2.
Novel three‐residue helix‐turn secondary structures, nucleated by a helix at the N terminus, were generated in peptides that have ‘β‐Caa‐L ‐Ala‐L ‐Ala,’ ‘β‐Caa‐L ‐Ala‐γ‐Caa,’ and ‘β‐Caa‐L ‐Ala‐δ‐Caa’ (in which βCaa is C‐linked carbo‐β‐amino acid, γCaa is C‐linked carbo‐γ‐amino acid, and δ‐Caa is C‐linked carbo‐δ‐amino acid) at the C terminus. These turn structures are stabilized by 12‐, 14‐, and 15‐membered (mr) hydrogen bonding between NH(i)/CO(i+2) (i+2 is the last residue in the peptide) along with a 7‐mr hydrogen bond between CO(i)/NH(i+2). In addition, a series of α/β‐peptides were designed and synthesized with alternating glycine (Gly) and (S)‐β‐Caa to study the influence of an achiral α‐residue on the helix and helix‐turn structures. In contrast to previous results, the three ‘β–α–β’ residues at the C terminus (α‐residue being Gly) are stabilized by only a 13‐mr forward hydrogen bond, which resembles an α‐turn. Extensive NMR spectroscopic and molecular dynamics (MD) studies were performed to support these observations. The influence of chirality and side chain is also discussed.  相似文献   

3.
Pasteur’s major discovery in chemistry was the recognition of molecular chirality, in 1848. He understood that his new science needed its own language, and introduced new terminology and nomenclature, thereby launching modern stereochemical language. He was eminently prepared for this task as a refined user of language, skills recognized by his election to the Académie française, the supreme institution for the protection and promotion of the French language. The terms chiral and chirality did not exist at the time and he adopted the French word dissymmétrie (dissymmetry) for the phenomenon of handedness. Although in his time almost nothing was known about molecular constitution and configuration, his insights allowed him to create useful language some of which is still used today in stereochemistry, e. g., racemic for the 1 : 1 mixture of the two enantiomers, and the use of the prefixes levo‐ and dextro‐ in the names of optically active substances. On the other hand, the limitations in the knowledge of organic chemistry at the time prevented him from creating some needed terms, e. g., for the phenomenon of diastereoisomerism. He also failed to adopt the enantio terminology introduced in the 1850s by German mineralogist Carl Friedrich Naumann. Analysis of Pasteur’s linguistic innovations is of interest from the point of view of the history of chemistry and is also useful in throwing light on the fundamental nature of the concepts of stereochemistry. Such understanding has acquired a new relevance due to the considerable misuse and misunderstanding of this language seen in the literature today.  相似文献   

4.
The syntheses of various strapped and ‘picket‐fence’ chiral porphyrins are described, and their reactivities towards the enantioselective epoxidation of alkenes are reported. Four L ‐proline residues provide the chirality for the various meso‐substituted catalysts, which differ by either the spatial arrangement of the stereogenic centers or the nature and length of the straps. The resulting bridged structures possess four amide linkages in each strap, leading to highly rigid molecules with well‐defined geometries whereas the strapped Fe catalysts gave rise to only moderate enantioselectivities, the C2‐symmetrical ones being superior to the D2‐symmetrical compounds. The D2‐symmetrical ‘picket‐fence’ porphyrins were as selective as their strapped counterparts.  相似文献   

5.
Four novel chiral phenylacetylenes having an L ‐amino alcohol residue and two hydroxymethyl groups were synthesized and polymerized by an achiral catalyst ((nbd)Rh+6‐(C6H5)B?(C6H5)3]) or a chiral catalytic system ([Rh(nbd)Cl]2/(S)‐ or (R)‐phenylethylamine ((S)‐ or (R)‐PEA)). The two resulting polymers having an L ‐valinol or L ‐phenylalaninol residue showed Cotton effects at wavelengths around 430 nm. This observation indicated that they had an excess of one‐handed helical backbones. Positive and negative Cotton effects were observed only for the polymers having an L ‐valinol residue produced by using (R)‐ and (S)‐PEA as a cocatalyst, respectively, although the monomer had the same chirality. Even when the achiral catalyst was used, the two resulting polymers having an L ‐valinol or L ‐phenylalaninol residue showed Cotton effects despite the long distance between the chiral groups and the main chain. We have found the first example of a new type of chiral monomer, that is, a chiral phenylacetylene monomer having an L ‐amino alcohol residue and two hydroxy groups that was suitable for both modes of asymmetric polymerization, that is, the helix‐sense‐selective polymerization ( HSSP ) with the chiral catalytic system and the asymmetric‐induced polymerization ( AIP ) with the achiral catalyst. The other two monomers having L ‐alaninol and L ‐tyrosinol were found to be unsuitable to neither HSSP nor AIP because of their polymers' low solubility. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012  相似文献   

6.
In the context of Eschenmoser's work on pyranosyl‐RNA (‘p‐RNA’), we investigated the synthesis and base‐pairing properties of the 5‐methylisocytidine derivative. The previously determined clear‐cut restrictions of base‐pairing modes of p‐RNA had led to the expectation that a 5‐methylisocytosine β‐D ‐ribopyranosyl (= D ‐pr(MeisoC)) based (4′ → 2′)‐oligonucleotide would pair inter alia with D ‐pr(isoG) and L ‐pr(G) based oligonucleotides (D ‐pr and L ‐pr = pyranose form of D ‐ and L ‐ribose, resp.). Remarkably, we could not observe pairing with the D ‐pr(isoG) oligonucleotide but only with the L ‐pr(G) oligonucleotide. Our interpretation concludes that this – at first hand surprising – observation is caused by a change in the nucleosidic torsion angle specific for isoC.  相似文献   

7.
Novel 4‐ethynylphthaloyl amino acid esters carrying different terminal groups, 4‐ethynylphthaloyl glycine (1S,2R,5S)‐menthyl ester ( 1 ), 4‐ethynylphthaloyl glycine (1R,2S,5R)‐menthyl ester ( 2 ), 4‐ethynylphthaloyl L ‐leucine methyl ester ( 3 ), 4‐ethynylphthaloyl L ‐leucine (1S,2R,5S)‐menthyl ester ( 4 ), 4‐ethynylphthaloyl L ‐leucine (1R,2S,5R)‐menthyl ester ( 5 ) were synthesized and polymerized with a rhodium catalyst. Polymers with high molecular weights were obtained in 71–92% yields. The helical conformation of the polymers could be tuned by the chirality of the amino acid connected to the backbone, together with the chirality and bulkiness of the terminal pendent groups. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4183–4192, 2008  相似文献   

8.
《化学:亚洲杂志》2017,12(15):1839-1850
Chiral nanomaterials have received wide interest in many areas, but the exact origin of chirality at the atomic level remains elusive in many cases. With recent significant progress in atomically precise gold nanoclusters (e.g., thiolate‐protected Aun (SR)m ), several origins of chirality have been unveiled based upon atomic structures determined by using single‐crystal X‐ray crystallography. The reported chiral Aun (SR)m structures explicitly reveal a predominant origin of chirality that arises from the Au–S chiral patterns at the metal–ligand interface, as opposed to the chiral arrangement of metal atoms in the inner core (i.e. kernel). In addition, chirality can also be introduced by a chiral ligand, manifested in the circular dichroism response from metal‐based electronic transitions other than the ligand's own transition(s). Lastly, the chiral arrangement of carbon tails of the ligands has also been discovered in a very recent work on chiral Au133(SR)52 and Au246(SR)80 nanoclusters. Overall, the origins of chirality discovered in Aun (SR)m nanoclusters may provide models for the understanding of chirality origins in other types of nanomaterials and also constitute the basis for the development of various applications of chiral nanoparticles.  相似文献   

9.
A series of quadruple‐stranded Na+ and Ca2+ complexes with octadentate cyclen ligands was synthesized to produce complexes that contained four different side‐arm combinations (one triazole? coumarin group and three pyridine groups ( 1 ), four pyridine groups ( 2 ), one triazole? coumarin group and three quinoline groups ( 3 ), and four quinoline groups ( 4 )). X‐ray crystallographic analysis revealed that no significant changes occurred in the stereostructure of these complexes upon replacing one pyridine group with a triazole? coumarin moiety, or by replacing Na+ ions with Ca2+ ions, although the coordination number of the complexes in the solid state decreased when pyridine groups were replaced by quinoline groups. In solution, all of the side arms were arranged in a propeller‐like pattern to yield an enantiomer pair of Δ and Λ forms in each metal complex. The addition of a tert‐butoxycarbonyl (Boc)‐protected amino acid anion, that is, a coordinative chiral carboxylate anion, to the cyclen? Ca2+ complex induced circular dichroism (CD) signals in the aromatic region by forming a 1:1 mixture of diastereomeric ternary complexes with opposite complex chirality, whilst the corresponding Na+ complexes rarely showed any response. In complexes 1 ‐Ca2+ and 3 ‐Ca2+, this chirality‐transfer process was efficiently followed by considering the induction of the CD signals at two different wavelengths, that is, the coumarin‐chromophore region and the aza‐aromatic region. The sign and intensity of the CD signal were significantly dependent on both the nature of the aza‐aromatic moiety and the enantiomeric purity of the external anion. These Ca2+ complexes worked as effective probes for the determination of the enantiomeric excess of the chiral anion. The cyclen? Ca2+ complexes also interacted with the non‐coordinative Δ‐TRISPHAT anion through an ion‐pairing mechanism to achieve chirality transfer from the anion to the metal complex; both complexes 1 ‐Ca2+ and 3 ‐Ca2+ clearly showed induced CD signals in the coumarin‐chromophore region, owing to ion‐paring interactions with the Δ‐TRISPHAT anion. Thus, the proper combination of an octadentate cyclen ligand and a metal center demonstrated effective chirality transfer.  相似文献   

10.
Two new series of Boc‐N‐α,δ‐/δ,α‐ and β,δ‐/δ,β‐hybrid peptides containing repeats of L ‐Ala‐δ5‐Caa/δ5‐Caa‐L ‐Ala and β3‐Caa‐δ5‐Caa/δ5‐Caa‐β3‐Caa (L ‐Ala = L ‐alanine, Caa = C‐linked carbo amino acid derived from D ‐xylose) have been differentiated by both positive and negative ion electrospray ionization (ESI) ion trap tandem mass spectrometry (MS/MS). MSn spectra of protonated isomeric peptides produce characteristic fragmentation involving the peptide backbone, the Boc‐group, and the side chain. The dipeptide positional isomers are differentiated by the collision‐induced dissociation (CID) of the protonated peptides. The loss of 2‐methylprop‐1‐ene is more pronounced for Boc‐NH‐L ‐Ala‐δ‐Caa‐OCH3 (1), whereas it is totally absent for its positional isomer Boc‐NH‐δ‐Caa‐L ‐Ala‐OCH3 (7), instead it shows significant loss of t‐butanol. On the other hand, second isomeric pair shows significant loss of t‐butanol and loss of acetone for Boc‐NH‐δ‐Caa‐β‐Caa‐OCH3 (18), whereas these are insignificant for its positional isomer Boc‐NH‐β‐Caa‐δ‐Caa‐OCH3 (13). The tetra‐ and hexapeptide positional isomers also show significant differences in MS2 and MS3 CID spectra. It is observed that ‘b’ ions are abundant when oxazolone structures are formed through five‐membered cyclic transition state and cyclization process for larger ‘b’ ions led to its insignificant abundance. However, b1+ ion is formed in case of δ,α‐dipeptide that may have a six‐membered substituted piperidone ion structure. Furthermore, ESI negative ion MS/MS has also been found to be useful for differentiating these isomeric peptide acids. Thus, the results of MS/MS of pairs of di‐, tetra‐, and hexapeptide positional isomers provide peptide sequencing information and distinguish the positional isomers. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

11.
The systemic investigation of the structural impacts of side chains on the pH‐ and thermo‐responsiveness of tertiary amine functionalized poly(l ‐glutamate)s (TA‐PGs) was carried out. The TA‐PGs polymers were effectively synthesized by Cu(I)‐catalyzed azide‐alkyne cycloaddition click reaction of azido tertiary amines with poly(γ‐propargyl‐l ‐glutamate) (PPLG). Turbimetric measurements were performed to characterize the pH‐ and temperature‐induced phase transition of TA‐PGs in aqueous solution, which suggested a structural dependence of the properties on the N‐substituted groups and the “linkers” between 1,2,3‐triazole ring and the tertiary amine groups in the side chains. In detail, the pH responsive properties of TA‐PGs were basically determined by the hydrophobicity of the N‐substituted groups in the side chains and the pH transition point (pHt) decreased as the increasing hydrophobicity of the N‐substituted groups, while the temperature‐responsiveness of TA‐PGs were affected by either the N‐substituted groups or the “linkers.” TA‐PGs with a moderate N‐substituted amine group (e.g., DEA, PR, and PD) or a branched “linker” (e.g., iso‐propylene and 2‐methylpropylene group) were more likely to express the LCST‐type phase transition tuned by pH variation. These structure–property relationships revealed in this study would help to develop the applications of TA‐PGs in smart drug delivery systems. © 2013 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 671–679  相似文献   

12.
A new kind of ‘para‐effect’ under electron ionization (EI) conditions has been discovered for a series of bis(perfluoroacyl) derivatives of o‐, m‐ and p‐phenylenediamines, ‐hydroxybenzeneamines and ‐mercaptobenzeneamines of a common structure RCOX–C6H4–NHCOR (X = NH, S, O; R = CF3, C2F5, C3F7). Only the para‐isomers showed successive loss of a radical RCO? and a molecule RCN, leading to very intense peaks in the EI spectra. The composition and the origin of the [M–COR–NCR]+ ions were confirmed by exact mass measurements and linked scan experiments. The proposed mechanism of their formation takes into account likely para‐quinoid structures of the precursor ions. A similar rearrangement has not been observed for para‐isomers in the series of bis(perfluoroacyl) derivatives of benzenediols, mercaptophenols and dimercaptobenzenes. Published in 2010 by John Wiley & Sons, Ltd.  相似文献   

13.
(+)‐(1S)‐1,1′‐Binaphthalene‐2,2′‐diyl hydrogen phosphate (bnppa) is one of the useful optical selectors. To disclose the molecular mechanism by which bnppa recognizes aliphatic L ‐α‐amino acids and separates them by fractional crystallization, X‐ray analyses of bnppa and of its salts with L ‐alanine, L ‐valine, L ‐norvaline, and L ‐norleucine have been undertaken. All the amino acids adopt energetically favorable conformations in the crystal structures. The conformations and the packing patterns of bnppa in these crystal structures are very similar. The bnppa molecules are packed in a specific way to form hydrophobic and hydrophilic layers that are well separated. Between bnppa molecules, at the interface of these hydrophobic and hydrophilic layers, a space with chirality is formed. This space, designated as chiral space, recognizes the optically active amino acids. The packing of bnppa is mainly governed by intermolecular CH⋅⋅⋅π interactions between naphthalene moieties. The chiral space is responsible for the molecular recognition by bnppa allowing fractional crystallization of the L ‐α‐amino acids.  相似文献   

14.
Spontaneous mirror‐symmetry breaking is a fundamental process for development of chirality in natural and in artificial self‐assembled systems. A series of triple chain azobenzene based rod‐like compounds is investigated that show mirror‐symmetry breaking in an isotropic liquid occurring adjacent to a lamellar LC phase. The transition between the lamellar phase and the symmetry‐broken liquid is affected by trans cis photoisomerization, which allows a fast and reversible photoinduced switching between chiral and achiral states with non‐polarized light.  相似文献   

15.
The syntheses of phenacyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate and allyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate are reported. Reactions of these 2H‐azirin‐3‐amine derivatives with Z‐protected amino acids have shown them to be suitable synthons for the Aib‐Pro unit in peptide synthesis. After incorporation into the peptide by means of the ‘azirine/oxazolone method’, the C‐termini of the resulting peptides were deprotected selectively with Zn in AcOH or by a mild Pd0‐promoted procedure, respectively.  相似文献   

16.
Separation of racemic mixture of (RS)‐bupropion, (RS)‐baclofen and (RS)‐etodolac, commonly marketed racemic drugs, has been achieved by modifying the conventional ligand exchange approach. The Cu(II) complexes were first prepared with a few l ‐amino acids, namely, l ‐proline, l ‐histidine, l ‐phenylalanine and l ‐tryptophan, and to these was introduced a mixture of the enantiomer pair of (RS)‐bupropion, or (RS)‐baclofen or (RS)‐etodolac. As a result, formation of a pair of diastereomeric complexes occurred by ‘chiral ligand exchange’ via the competition between the chelating l ‐amino acid and each of the two enantiomers from a given pair. The diastereomeric mixture formed in the pre‐column process was loaded onto HPLC column. Thus, both the phases during chromatographic separation process were achiral (i.e. neither the stationary phase had any chiral structural feature of its own nor did the mobile phase have any chiral additive). Separation of diastereomers was successful using a C18 column and a binary mixture of MeCN and TEAP buffer of pH 4.0 (60:40, v/v) as mobile phase at a flow rate of 1 mL/min and UV detection at 230 nm for (RS)‐Bup, 220 nm for (RS)‐Bac and 223 nm for (RS)‐Etd. Baseline separation of the two enantiomers was obtained with a resolution of 6.63 in <15 min. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

17.
Two perylene diimide (PDI) enantiomers ( d/l ‐PDI ) incorporating the d /l ‐alanine moiety have been designed and synthesized. d/l ‐PDI in chloroform displays bright‐yellow fluorescence that is redshifted to orange‐red when the solvent contains a methanol fraction of 99 vol %. No circular dichroism (CD) or circularly polarized luminescence (CPL) signals were observed for d/l ‐PDI enantiomers in CHCl3. Interestingly, the d/l ‐PDI enantiomers exhibit clear mirror‐image Cotton effects and CPL emission in the aggregate state. The optical anisotropy factor (glum) is as high as 0.02 at fm=99 %, which can be attributed to self‐assembly through intermolecular π–π interactions in the aggregate state.  相似文献   

18.
The second leading cause of death in the US is cancer and early discovery of the disease has translated into reduced fatality rates. We have identified and performed a systematic investigation of a method for urinary pteridine analysis by using CE‐LIF, which is believed to possess the potential to diagnose the presence of cancer even earlier than existing methodologies. Through system enhancements, we have been able to improve the resolution of the two least resolved sets of peaks (6,7‐dimethylpterin versus 6‐biopterin and D ‐(+)‐neopterin versus 6‐hydroxymethylpterin) from 0.85 to 2.48 and 0.90 to 3.58, respectively. Additionally, we have discovered that the preparation of the urine samples in previous works was inadequate, and we have corrected the method to fully oxidize the pteridines in the urine, resulting in significantly less variability in quantification and greater ease of defining p‐values for healthy versus cancer patients. Finally, we have performed validation steps of spike and recovery and short‐term aging studies to demonstrate the method's robustness. As a result, we present an optimized and validated method ready for transfer from discovery phase to clinical trial that can potentially act as a non‐invasively pre‐screening test for cancer.  相似文献   

19.
Supramolecular hidden chirality of hydrogen‐bonded (HB) networks of primary ammonium carboxylates was exposed by advanced graph set analysis from a symmetric viewpoint in topology. The ring‐type HB (R‐HB) networks are topologically regarded as faces, and therefore exhibit prochirality and positional isomerism due to substituents attached on the faces. To describe the symmetric properties of the faces, additional symbols, Re (right‐handed or clockwise), Si (left‐handed or anticlockwise), and m (mirror), were proposed. According to the symbols, various kinds of faces were classified based on the symmetry. This symmetry consideration of the faces enables us to precisely evaluate supramolecular chirality, especially its handedness, of 0D‐cubic, 1D‐ladder and 2D‐sheet HB networks that are composed of the faces. The 1D‐ladder and 2D‐sheet HB networks generate chirality by accumulating the chiral faces in 1D and 2D manners, respectively, whereas 0D‐cubic HB networks generate chirality based on combinations of eight kinds of faces, similar to the chirality of dice.  相似文献   

20.
Mining microbial genomes including those of Streptomyces reveals the presence of a large number of biosynthetic gene clusters. Unraveling this genetic potential has proved to be a useful approach for novel compound discovery. Here, we report the heterologous expression of two similar P450‐associated cyclodipeptide synthase‐containing gene clusters in Streptomyces coelicolor and identification of eight rare and novel natural products, the C3‐guaninyl indole alkaloids guanitrypmycins. Expression of different gene combinations proved that the cyclodipeptide synthases assemble cyclo‐l ‐Trp‐l ‐Phe and cyclo‐l ‐Trp‐l ‐Tyr, which are consecutively and regiospecifically modified by cyclodipeptide oxidases, cytochrome P450 enzymes, and N‐methyltransferases. In vivo and in vitro results proved that the P450 enzymes function as key biocatalysts and catalyze the regio‐ and stereospecific 3α‐guaninylation at the indole ring of the tryptophanyl moiety. Isotope‐exchange experiments provided evidence for the non‐enzymatic epimerization of the biosynthetic pathway products via keto–enol tautomerism. This post‐pathway modification during cultivation further increases the structural diversity of guanitrypmycins.  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