首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
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   

2.
Natural polynucleotides including Micrococcus lysodeikticus and calf thymus DNA exhibit enantioselective recognition to S‐ofloxacin regulated by Cu2+. This is the first report that ofloxacin and Cu2+ have cooperative effects on the local distortions of polynucleotides. At the [Cu2+]/[base] ratio of 0.1, S‐ofloxacin is more liable to induce the locally distorted structures of polynucleotides, of which the association constant of S‐ofloxacin toward DNA‐Cu(II) is three times higher than that of the R‐enantiomer. The apparent increase of adsorption capability and cooperativity, as well as the change of adsorption mechanism were detected in the adsorption of ofloxacin enantiomers on polynucleotides upon Cu(II)‐coordination. This study not only discloses the effect of the chiral drug on the structural transition of long double‐stranded DNA, but provides fundamental data to develop a novel enantioseparation method based on natural polynucleotides. Chirality 27:306‐313, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

3.
To investigate the structural role played by isostructural unbranched alkyl‐chains on the conformational ensemble and stability of β‐turn structures, the conformational properties of a designed model peptide: Plm‐Pro‐Gly‐Pda ( 1 , Plm: H3C—(CH2)14—CONH—; Pda: —CONH— (CH2)14—CH3) have been examined and compared with the parent peptide: Boc‐Pro‐Gly‐NHMe ( 2 , Boc: tert‐butoxycarbonyl; NHMe: N‐methylamide). The characteristic 13C NMR chemical‐shifts of the Pro Cβ and Cγ resonances ascertained the incidence of an all‐trans peptide‐bond in low polarity deuterochloroform solution. Using FTIR and 1H NMR spectroscopy, we establish that apolar alkyl‐chains flanking a β‐turn promoting Pro‐Gly sequence impart definite incremental stability to the well‐defined hydrogen‐bonded structure. The assessment of 1H NMR derived thermodynamic parameters of the hydrogen‐bonded amide‐NHs via variable temperature indicate that much weaker hydrophobic interactions do contribute to the stability of folded reverse turn structures. The far‐UV CD spectral patterns of 1 and 2 in 2,2,2‐trifluoroethanol are consistent with Pro‐Gly specific type II β‐turn structure, concomitantly substantiate that the flanking alkyl‐chains induce substantial bias in enhanced β‐turn populations. In view of structural as well as functional importance of the Pro‐Gly mediated secondary structures, besides biochemical and biological significance of proteins lipidation via myristoylation or palmytoilation, we highlight potential convenience of the unbranched Plm and Pda moieities not only as main‐chain N‐ and C‐terminal protecting groups but also to mimic and stabilize specific isolated secondary and supersecondary structural components frequently observed in proteins and polypeptides. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 419–426, 2013.  相似文献   

4.
DNA‐based chiral selectors are constructed to discriminate ofloxacin enantiomers through metal‐ion anchoring on a special DNA double helix that contains successive GC pairs. The effects of metal ions involving Mg2+, Ni2+, Cu2+, Ag+, and Pt2+ were studied on the regulation of DNA chiral discrimination towards ofloxacin enantiomers. It is shown that DNA‐Cu(II) complexes exhibit the highest enantioselectivities at the [Cu2+]/base ratio of 0.1. The enantiomeric excess can reach 59% in R‐enantiomer after being adsorbed by the RET‐Cu(II) complex. Stereoselective recognition of ofloxacin enantiomers on the double helix is tunable via external stimulus, providing a programmable desorption process to regenerate DNA. This DNA‐based chiral selector exhibits excellent reusability without apparent loss of enantioselectivity after three cycles of adsorption and desorption. Chirality 26:249–254, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

5.
In this study, R(+)‐α‐methylbenzylamine‐modified magnetic chiral sorbent was synthesized and assessed as a new enantioselective solid phase sorbent for separation of mandelic acid enantiomers from aqueous solutions. The chemical structures and magnetic properties of the new sorbent were characterized by vibrating sample magnetometry, transmission electron microscopy, Fourier transform infrared spectroscopy, and dynamic light scattering. The effects of different variables such as the initial concentration of racemic mandelic acid, dosage of sorbent, and contact time upon sorption characteristics of mandelic acid enantiomers on magnetic chiral sorbent were investigated. The sorption of mandelic acid enantiomers followed a pseudo‐second‐order reaction and equilibrium experiments were well fitted to a Langmuir isotherm model. The maximum adsorption capacity of racemic mandelic acid on to the magnetic chiral sorbent was found to be 405 mg g?1. The magnetic chiral sorbent has a greater affinity for (S)‐(+)‐mandelic acid compared to (R)‐(?)‐mandelic acid. The optimum resolution was achieved with 10 mL 30 mM of racemic mandelic acid and 110 mg of magnetic chiral sorbent. The best percent enantiomeric excess values (up to 64%) were obtained by use of a chiralpak AD‐H column. Chirality 27:835–842, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

6.
The Zn(II)–cyclen–dipeptide ternary systems (where cyclen is abbreviated as L and dipeptide is glycylglycine (HL1) or glycyl‐(S)‐alanine (HL2)) were investigated by potentiometry applying both “out‐of‐cell” and direct titrations and by 1H NMR spectroscopy. Especially, the 1H NMR study was found to be very efficient to estimate speciation in the systems. The results obtained under full equilibria indicated two main species, [Zn(L)(HL1,2)]2+ and [Zn(L)(L1,2)]+, in both the systems. In the [Zn(L)(HL1,2)]2+ complex, presence of carbonyl‐carboxylate chelate was confirmed, and in the [Zn(L)(L1,2)]+ species, the peptide coordination is re‐organized to carbonyl‐amine chelate or only terminal amino group is coordinated. Equilibrium constants describing [Zn(L)]2+–dipeptide interaction are relatively low, log K = 3.4 for Gly‐Gly and 4.1 for Gly‐(S)‐Ala, respectively. Nevertheless, the values are slightly higher than stability constants for interaction of Zn(II) with the dipeptides (i.e. [Zn(L1,2)]+ species) where a chelate formation is expected. It indicates that interaction between Zn(II) ion in [Zn(L)]2+ and the dipeptides should be supported by some additional interactions. Potentiometry carried out under non‐equilibrum condition showed different species where these additional stabilizing forces play more important role. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

7.
Patterning of solid surfaces with functional organic molecules has been a convenient route to fabricate two‐dimensional materials with programmed architecture and activities. One example is the chiral nanoporous networks that can be created via controlled self‐assembly of star‐shaped molecules under 2D confinement. In this contribution we use computer modeling to predict the formation of molecular networks in adsorbed overlayers comprising cruciform molecular building blocks equipped with discrete interaction centers. To that end, we employ the Monte Carlo simulation method combined with a coarse‐grained representation of the adsorbed molecules which are treated as collections of interconnected segments. The interaction centers within the molecules are represented by active segments whose number and distribution are adjusted. Our particular focus is on those distributions that produce prochiral molecules able to occur in adsorbed configurations being mirror images of each other (surface enantiomers). We demonstrate that, depending on size, aspect ratio, and intramolecular distribution of active sites, the surface enantiomers can co‐crystallize or segregate into extended homochiral domains with largely diversified nanosized cavities. The insights from our theoretical studies can be helpful in designing 2D chiral porous networks with potential applications in enantioselective adsorption and asymmetric heterogeneous catalysis. Chirality 27:397–404, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

8.
Two new chiral mononuclear Mn(III) complexes, [Mn L ( R )Cl (C2H5OH)]?C2H5OH ( 1 ) and [Mn L ( S ) (CH3OH)2]Cl?CH3OH ( 2 ), {H2 L = (R,R)‐or (S,S)‐N,N’‐bis‐(2‐hydroxy‐1‐naphthalidehydene)‐cyclohexanediamine} were synthesized and characterized by various physicochemical techniques. Bond valence sum (BVS) calculations and the Jahn‐Teller effect indicate that the Mn centers are in a +3 oxidation state. The statuses of the two complexes in the solution were confirmed as a pair of enantiomers by electrospray ionization, mass spectrometry (ESI‐MS) spectrum. The binding ability of the complexes with calf thymus CT‐DNA was investigated by spectroscopic and viscosity measurements. Both of the complexes could interact with CT‐DNA via an intercalative mode with the order of 1 ( R ‐enantiomer) > 2 ( S ‐enantiomer). Under the physiological conditions, the two compounds exhibit efficient DNA cleavage activities without any external agent, which also follows the order of R ‐enantiomer > S ‐enantiomer. Interestingly, the concentration‐dependent DNA cleavage experiments indicate an optimal concentration of 17.5 μM. In addition, the interaction of the compounds with bovine serum albumin (BSA) was also investigated, which indicated that the complexes could quench the intrinsic fluorescence of BSA by a static quenching mechanism. Chirality 27:142‐150, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

9.
A series of terminally blocked peptides (to the pentamer level) from l ‐Ala and the cyclic Cα,α‐disubstituted Gly residue Afc and one Gly/Afc dipeptide have been synthesized by solution method and fully characterized. The molecular structure of the amino acid derivative Boc‐Afc‐OMe and the dipeptide Boc‐Afc‐Gly‐OMe were determined in the crystal state by X‐ray diffraction. In addition, the preferred conformation of all of the model peptides was assessed in deuterochloroform solution by FT‐IR absorption and 1H‐NMR. The experimental data favour the conclusion that the Afc residue tends to adopt either the fully‐extended (C5) or a folded/helical structure. In particular, the former conformation is highly populated in solution and is also that found in the crystal state in the two compounds investigated. A comparison with the structural propensities of the strictly related Cα,α‐disubstituted Gly residues Ac5c and Dϕg is made and the implications for the use of the Afc residue in conformationally constrained analogues of bioactive peptides are briefly examined. A spectroscopic (UV absorption, fluorescence, CD) characterization of this novel aromatic Cα,α‐disubstituted Gly residue is also reported. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Chiral ionic liquids (CILs) with amino acids as cations have been applied as novel chiral ligands coordinated with Cu2+ to separate tryptophan enantiomers in ligand exchange chromatography. Four kinds of amino acid ionic liquids, including [L‐Pro][CF3COO], [L‐Pro][NO3], [L‐Pro]2[SO4], and [L‐Phe][CF3COO] were successfully synthesized and used for separation of tryptophan enantiomers. To optimize the separation conditions, [L‐Pro][CF3COO] was selected as the model ligand. Some factors influencing the efficiency of chiral separation, such as copper ion concentration, CILs concentration, methanol ratio (methanol/H2O, v/v), and pH, were investigated. The obtained optimal separation conditions were as follows: 8.0 mmol/L Cu(OAc)2, 4.0 mmol/L [L‐Pro][CF3COO] ,and 20% (v/v) methanol at pH 3.6. Under the optimum conditions, acceptable enantioseparation of tryptophan enantiomers could be observed with a resolution of 1.89. The results demonstrate the good applicability of CILs with amino acids as cations for chiral separation. Furthermore, a comparative study was also conducted for exploring the mechanism of the CILs as new ligands in ligand exchange chromatography. Chirality 26:160–165, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

11.
We have recorded high-resolution 13C-nmr spectra of collagen fibrils in the solid state by the cross-polarization–magic-angle-spinning(CP–MAS)method and analyzed the spectra with reference to those of collagenlike polypeptides. We used two kinds of model polypeptides to obtain reference 13C chemical shifts of major amino acid residues of collagen (Gly, Pro, Ala, and Hyp): the 31-helical polypeptides [(Gly)nII, (Pro)nII, (Hyp)n, and (Ala? Gly? Gly)nII], and the triple-helical polypeptides [(Pro? Gly? Pro)n and (Pro? Ala? Gly)n]. Examination of the 13C chemical shifts of these polypeptides, together with our previous data, showed that the 13C chemical shifts of individual amino acid residues are the same, within experimental error (±0.5 ppm), among different polypeptides with different primary sequences, if the conformations are the same. We found that the 13C chemical shifts of Ala residues of the 31-helical (Ala? Gly? Gly)n and triple-helical (Pro? Ala? Gly)n are significantly displaced, compared with those of the α-helix, β-sheet, and silk I form, and can be utilized as excellent probes to examine conformational features of collagen-like polypeptides. Further, the 13C chemical shifts of Gly and Pro residues in the triple-helical polypeptides are substantially displaced from those found in (Gly)nII and (Pro)nII of the 31-helix, reflecting further conformational change from the 31-helix to the supercoiled triple helix. In particular, the 13C chemical shifts of Gly C ? O carbons of the triple-helical polypeptides are substantially displaced upfield (4.1–5.1 ppm), with respect to those of the 31-helical polypeptides. These displacements are interpreted by that Gly C ? O of the former is not involved in NH …? O ? C hydrogen bonds, while this carbon of the latter is linked by these kinds of hydrogen bonds. On the basis of these 13C chemical shifts, as reference data for the collagenlike structure, we were able to assign the 13C-nmr peaks of Gly, Ala, Pro, and Hyp residues of collagen fibrils, which are in good agreement with the values expected from the model polypeptides mentioned above. We also discuss a plausible conformational change of collagen fibrils during denaturation.  相似文献   

12.
Chiral ionic liquids hold promise in many asymmetric applications. This study explores the impact of ionic solutes on the chiral discrimination of five amino acid methyl ester‐based ionic liquids, including L‐ and D‐alanine methyl ester, L‐proline methyl ester, L‐leucine methyl ester, and L‐valine methyl ester cations combined with bis(trifluoromethanesulfonimide) anion. Circularly polarized luminescence spectroscopy was used to study the chiral discrimination by measuring the racemization equilibrium of a dissymmetric europium complex, Eu(dpa)33? (where dpa = 2,6‐pyridinedicarboxylate). The chiral discrimination measured was dependent on the concentration of Eu(dpa)33? and this concentration‐dependence was different in each of the ionic liquids. Ionic liquids with L‐leucine methyl ester and L‐valine methyl ester even switched enantiomeric preference based on the solute concentration. Changing the cation of the Eu(dpa)33? salt from tetrabutylammonium to tetramethylammonium ion also affected the chiral discrimination demonstrated by the ionic liquids. Chirality 27:320–325, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
A practical synthetic method for 2,2'‐disubstituted fluorinated binaphthyl derivatives was achieved using magnesium bis(2,2,6,6‐tetramethylpiperamide) [Mg(TMP)2], prepared from LiTMP (2 equiv) and MgBr2 (1 equiv), which allows for access to a variety of fluorinated binaphthyl compounds. The utility of the fluorinated binaphthyl backbone was evaluated in F10BINOL derived chiral mono‐phosphoric acid (R)‐ 19 as the chiral Brønsted acid catalyst. The catalyst (R)‐ 19 performs exceptionally well in the catalytic enantioselective imino‐ene reaction, demonstrating the potential of a fluorinated binaphthyl framework. Chirality 27:464–475, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

14.
The synthesis of new dermorphin analogues is described. The (R)‐alanine or phenylalanine residues of natural dermorphin were substituted by the corresponding α‐methyl‐β‐azidoalanine or α‐benzyl‐β‐azido(1‐piperidinyl)alanine residues. The potency and selectivity of the new analogues were evaluated by a competitive receptor binding assay in rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). The most active analogue in this series, Tyr‐(R)‐Ala‐(R)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 and its epimer were analysed by 1H and 13C NMR spectroscopy and restrained molecular dynamics simulations. The dominant conformation of the investigated peptides depended on the absolute configuration around Cα in the α‐benzyl‐β‐azidoAla residue in position 3. The (R) configuration led to the formation of a type I β‐turn, whilst switching to the (S) configuration gave rise to an inverse β‐turn of type I′, followed by the formation of a very short β‐sheet. The selectivity of Tyr‐(R)‐Ala‐(R) and (S)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 was shown to be very similar; nevertheless, the two analogues exhibited different conformational preferences. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
The solid state conformations of cyclo[Gly–Proψ[CH2S]Gly–D –Phe–Pro] and cyclo[Gly–Proψ[CH2–(S)–SO]Gly–D –Phe–Pro] have been characterized by X-ray diffraction analysis. Crystals of the sulfide trihydrate are orthorhombic, P212121, with a = 10.156(3) Å, b = 11.704(3) Å, c = 21.913(4) Å, and Z = 4. Crystals of the sulfoxide are monoclinic, P21, with a = 10.662(1) Å, b = 8.552(3) Å, c = 12.947(2) Å, β = 94.28(2), and Z = 2. Unlike their all-amide parent, which adopts an all-trans backbone conformation and a type II β-turn encompassing Gly-Pro-Gly-D -Phe, both of these peptides contain a cis Gly1-Pro2 bond and form a novel turn structure, i.e., a type II′ β-turn consisting of Gly–D –Phe–Pro–Gly. The turn structure in each of these peptides is stabilized by an intramolecular H bond between the carbonyl oxygen of Gly1 and the amide proton of D -Phe4. In the cyclic sulfoxide, the sulfinyl group is not involved in H bonding despite its strong potential as a hydrogen-bond acceptor. The crystal structure made it possible to establish the absolute configuration of the sulfinyl group in this peptide. The two crystal structures also helped identify a type II′ β-turn in the DMSO-d6 solution conformers of these peptides. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
The photocatalytic reduction of nitrogen (N2) with water (H2O) as the reducing agent holds great promise as a sustainable future technology for the synthesis of ammonia (NH3). Herein, the effect of oxygen vacancies and electron‐rich Cuδ+ on the performance of zinc‐aluminium layered double hydroxide (ZnAl‐LDH) nanosheet photocatalysts for N2 reduction to NH3 under UV–vis excitation is systematically explored. Results show that a 0.5%‐ZnAl‐LDH nanosheet photocatalyst (containing 0.5 mol% Cu by metal basis) affords a remarkable NH3 production rate of 110 µmol g?1 h?1 and excellent stability in pure water. The X‐ray absorption spectroscopy, electron paramagnetic resonance, and density functional theory calculations reveal that Cu addition imparts oxygen vacancies and coordinatively unsaturated Cuδ+ (δ < 2) with electron‐rich property in the ZnAl‐LDH nanosheets, both of which readily contribute to efficient separation and transfer of photogenerated electrons and holes and promote N2 adsorption, thereby both activating N2 and facilitating its multielectrons reduction to NH3.  相似文献   

17.
The reduction of the axially chiral N‐(o‐aryl)‐5,5‐dimethyl‐2,4‐oxazolidinediones by NaBH4 yielded axially chiral N‐(o‐aryl)‐4‐hydroxy‐5,5‐dimethyl‐2‐oxazolidinone enantiomers having a chiral center at C‐4, with 100% diastereoselectivity as has been shown by their 1H and 13C NMR spectra and by enantioselective HPLC analysis. The resolved enantiomeric isomers were found to interconvert thermally through an aldehyde intermediate formed upon ring cleavage via a latent ring‐chain‐ring tautomerization. It was found that the rate of enantiomerization depended on the size and the electronic effect of the ortho substituent present on the aryl ring bonded to the nitrogen of the heterocycle. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

18.
The collagen triple helix has a larger accessible surface area per molecular mass than globular proteins, and therefore potentially more water interaction sites. The effect of deuterium oxide on the stability of collagen model peptides and Type I collagen molecules was analyzed by circular dichroism and differential scanning calorimetry. The transition temperatures (Tm) of the protonated peptide (Pro‐Pro‐Gly)10 were 25.4 and 28.7°C in H2O and D2O, respectively. The increase of the Tm of (Pro‐Pro‐Gly)10 measured calorimetrically at 1.0°C min?1 in a low pH solution from the protonated to the deuterated solvent was 5.1°C. The increases of the Tm for (Gly‐Pro‐4(R)Hyp)9 and pepsin‐extracted Type I collagen were measured as 4.2 and 2.2°C, respectively. These results indicated that the increase in the Tm in the presence of D2O is comparable to that of globular proteins, and much less than reported previously for collagen model peptides [Gough and Bhatnagar, J Biomol Struct Dyn 1999, 17, 481–491]. These experimental results suggest that the interaction of water molecules with collagen is similar to the interaction of water with globular proteins, when the ratio of collagen to water is very small and collagen is monomerically dispersed in the solvent. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 93–101, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

19.
Ming Wang  Wei Xie  Aijing Li  Sichuan Xu 《Chirality》2016,28(10):674-685
In order to investigate the chiral benzedrine molecules corresponding to their different characteristics in biochemical systems, we studied their interaction with D3R using the docking method, molecular dynamic simulation, and quantum chemistry. The obtained results indicate that the active residues for R‐benzedrine (RAT) bound with D3R are Ala132, Asp133, and Tyr55, while Asn57, Asp133, Asp168, Cys172, Gly54, Trp24, and Vall136 act as the active residues for S‐benzedrine (SAT). The different active pockets are observed for ART or SAT because they possess different active residues. The binding energies between RAT and SAT with D3R were determined to be ?44.0 kJ.mol?1 and ?71.2 kJ.mol?1, respectively. These results demonstrate that SAT within the studied pocket of D3R has a stronger capability of binding with D3R, while it is more feasible for RAT to leave from the interior positions of D3R. In addition, the results suggest that the D3R protein can recognize chiral benzedrine molecules and influence their different addictive and pharmacological effects in biochemical systems. Chirality 28:674–685, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

20.
《Chirality》2017,29(3-4):107-114
Conjugated homopolymer poly(9,9‐bis(3‐((S )‐2‐methylbutylpropanoate))fluorene) (PSF) with chiral pendants was synthesized and characterized. Dissolution experiments show that PSF is well dissolved in racemic limonene at high temperature and begins aggregating upon sequential cooling treatment. The corresponding assemblies were transferred to quartz plate by the spin‐coating method. Comparably, film casting from chloroform solution was also prepared. Upon annealing thermal treatments, these PSF films exhibited perfect mirror circular dichroism (CD) Cotton effects and dissymmetry ratios. Optical absorption spectroscopy (UV‐vis), CD, and fluorescence spectroscopy results reveal that chiral side chains successfully induced M ‐ and P ‐helical structures in aggregates and films, and this significant difference was ascribed to their differential supramolecular conformations.  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