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 15 毫秒
1.
The infrared (IR), vibrational circular dichroism (VCD), and electronic circular dichroism (ECD) spectra of short cationic sequential peptides (L ‐Lys‐L ‐Ala‐L ‐Ala)n (n = 1, 2, and 3) were measured over a range of temperatures (20–90 °C) in aqueous solution at near‐neutral pH values in order to investigate their solution conformations and thermally induced conformational changes. VCD spectra of all three oligopeptides measured in the amide I′ region indicate the presence of extended helical polyproline II (PPII)‐like conformation at room temperature. UV‐ECD spectra confirmed this conclusion. Thus, the oligopeptides adopt a PPII‐like conformation, independent of the length of the peptide chain. However, the optimized dihedral angles ? and ψ are within the range ?82 to ?107° and 143–154°, respectively, and differ from the canonical PPII values. At elevated temperatures, the observed intensity and bandshape variations in the VCD and ECD spectra show that the PPII‐like conformation of the Lys‐Ala‐Ala sequence is still preferred, being in equilibrium with an unordered conformer at near‐neutral pH values within the range of temperatures from 20 to 90 °C. This finding was obtained from analysis of the temperature‐dependent spectra using the singular value decomposition method. The study presents KAA‐containing oligopeptides as conformationally stable models of biologically important cationic peptides and proteins. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
We studied the interaction of poly‐l ‐lysine (PLL) and poly‐l ‐arginine (PLAG) with sodium dodecyl sulfate (SDS) surfactant and the interaction of poly‐l‐ glutamic acid (PLGA) and poly‐l ‐aspartic acid (PLAA) with tetradecyltrimethylammonium bromide (TTAB) surfactant using vibrational circular dichroism (VCD) spectroscopy in the region of C‐H stretching vibration and in the Amide I region both in solution and in mulls. A chirality transfer from polypeptides to achiral surfactants was observed in the C‐H stretching region, where measurements in solution were impossible. This observation was enabled by a special sample treatment technique using lyophilization and the preparation of mulls. This technique demonstrated itself as an interesting and beneficial tool for VCD measurements. In addition, we observed that SDS changed the secondary structure of PLL to the β‐sheet and of PLAG to the α‐helix. TTAB disrupted the PLGA and PLAA structure. These results were obtained in the mull but were confirmed by the VCD spectra measured in solution and by electronic circular dichroism. The chirality transfer from the polypeptides to SDS was caused by polypeptides ordered into a specific conformation during the interaction, while in the TTBA system it was induced primarily by the chirality of the amino acid residues. Chirality 27:965–972, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

3.
The increasing interest in peptidomimetics of biological relevance prompted us to synthesize a series of cyclic peptides comprising trans‐2‐aminocyclohexane carboxylic acid (Achc) or trans‐2‐aminocyclopentane carboxylic acid (Acpc). NMR experiments in combination with MD calculations were performed to investigate the three‐dimensional structure of the cyclic peptides. These data were compared to the conformational information obtained by electronic circular dichroism (ECD) and vibrational circular dichroism (VCD) spectroscopy. Experimental VCD spectra were compared to theoretical VCD spectra computed quantum chemically at B3LYP/6‐31G(d) density functional theory (DFT) level. The good agreement between the structural features derived from the VCD spectra and the NMR‐based structures underlines the applicability of VCD in studying the conformation of small cyclic peptides. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Noncovalent interactions of poly(L-lysine) (PL), oligopeptides L-lysyl-L-alanyl-L-alanine and (L-lysyl-L-alanyl-L-alanine)(2) with meso-tetrakis(4-sulfonatophenyl)porphine (TPPS), and poly(L-glutamic acid) (PLGA) with meso-tetrakis(1-methyl-4-pyridyl)porphine tetra-p-tosylate (TMPyP) in aqueous solutions have been studied using combination of spectroscopic methods: Vibrational circular dichroism (VCD) spectroscopy in the mid-infrared region provides a direct information on conformational changes of the polypeptides and oligopeptides caused by interactions with porphyrins; ultraviolet-visible absorption, fluorescence, and electronic circular dichroism (ECD) reveal the aggregation characterization of the porphyrin part of the complexes. Interactions of TPPS with tripeptide, hexapeptide, and PL containing about ten amino acid residues in the molecular chain are accompanied with the changes of VCD patterns in the amide I' region. In these cases, the conformation of the oligopeptide part of complexes is obviously influenced by interactions with TPPS and partial changes of random coil structure are observed in VCD. When PL was composed of the hundreds of lysine residues, just a weak intensity decrease was detected and the shape of VCD spectrum typical for the random coil structure was preserved. As follows from the uv-vis absorption and fluorescence spectra, porphyrin molecules are attached to peptides by electrostatic interaction as a monomer or dimer and interaction between porphyrin and peptide depends on the polypeptide chain length. For the PLGA-TMPyP system with PLGA containing from tens to hundreds of glutamic acid residues in the chain, the VCD spectra were unchanged when TMPyP was presented in the aqueous solution of PLGA and random coil conformation of PLGA-TMPyP aggregates was preserved.  相似文献   

5.
Two couples of enantiomeric platinum(II) complexes: Pt(L1a)Cl ( 1a ), Pt(L1b)Cl ( 1b ) and Pt(L1a)(C ≡ C ? Ph) ( 2a ), Pt(L1b)(C ≡ C ? Ph) ( 2b ) (L1a = (+)‐1,3‐di‐(2‐(4,5‐pinene)pyridyl)benzene, L1b = (?)‐1,3‐di‐(2‐(4,5‐pinene)pyridyl)benzene) were synthesized and characterized. Their absolute configurations were determined by single crystal X‐ray diffraction and further verified by circular dichroism (CD) spectra (including electronic circular dichroism [ECD] and vibrational circular dichroism [VCD]). These complexes show interesting mechanoluminescence and/or vapoluminescence due to crystalline‐to‐amorphous transformation. The crystalline solids, grinding‐induced amorphous powders, and vapor‐induced amorphous powders of complexes 2a and 2b were comparatively investigated by solid‐state ECD and VCD spectra. The transformation from crystalline solids to amorphous powders was accompanied by significant variances of the spectral feature in both ECD and VCD spectra. Chirality 25:384–392, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

6.
A series of representative optically active derivatives of 4‐hydroxy‐5‐alkylcyclopent‐2‐en‐1‐one were prepared from the respective 2‐furyl methyl carbinols via the Piancatelli rearrangement followed by the enzymatic kinetic resolution of racemates. Applicability of chiroptical methods (experimental and calculated electronic circular dichroism [ECD] and vibrational circular dichroism [VCD] spectra) to determine the absolute configuration of both stereogenic centers in 4‐hydroxy‐5‐methylcyclopent‐2‐en‐1‐one was demonstrated. It was also demonstrated that the concurrent application of ECD and VCD spectroscopy can be used for the determination of the configuration of two stereogenic centers. Chirality 26:300–306, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
The salen‐type ligand prepared with (R,R) diphenylethan‐1,2‐diamine and salicylaldehyde provides stable and inert complexes KLnL2 upon simple reaction with lanthanide halides or pseudohalides LnX3 (Ln = Tb3+‐Lu3+; X = Cl? or TfO?) of its potassium salt. All the complexes were completely characterized through nuclear magnetic resonance (NMR), electronic circular dichroism (ECD) in the UV and some (Er3+, Tm3+, Yb3+) also with Near‐IR ECD (NIR‐ECD) and luminescence (Tb3+, Tm3+). Careful analysis of the NMR shifts demonstrated that the complexes are isostructural in solution and afforded an accurate geometry. This was further confirmed by means of Density Functional Theory (DFT) optimization of the Lu3+ complex, and by comparing the ligand‐centered experimental and time‐dependent TD‐DFT computed UV‐ECD spectra. As final validation, we used the NIR‐ECD spectrum of the Yb3+ derivative calculated by means of Richardson's equations. The excellent match between calculated and experimental ECD spectra confirm the quality of the NMR structure.  Chirality 27:857–863, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

8.
The electronic circular dichroism (ECD) spectra of naproxen enantiomers were studied as a function of solvents using experimental (circular dichroism) and theoretical (time‐dependent density functional theory) approaches. The (R)‐ and (S)‐naproxen enantiomers presented an unusual inversion in their ECD signals in the presence of ethanol and water when compared with polar aprotic solvents such as acetonitrile. From a practical point of view, these findings deserve great attention because these solvents are widely used for high‐performance liquid chromatography analysis in quality control of chiral pharmaceutical drugs. This is particularly relevant to naproxen because the (S)‐naproxen has anti‐inflammatory properties, whereas (R)‐naproxen is hepatotoxic. A time‐dependent density functional theory computer simulation was conducted to investigate the signal inversion using the solvation model based on density, a reparameterization of polarized continuum model. Electronic circular dichroism signals of conformers were calculated by computer simulation and their contribution to the combined spectra obtained according to Boltzmann weighting. It was found that the experimentally observed ECD signal inversion can be associated with the minor or major contribution of different conformers of naproxen.  相似文献   

9.
The interactions were studied of DNA with the nonhistone chromatin protein HMGB1 and histone H1 in the presence of manganese(II) ions at different protein to DNA and manganese to DNA phosphate ratios by using absorption and optical activity spectroscopy in the electronic [ultraviolet (UV) and electronic circular dichroism ECD)] and vibrational [infrared (IR) and vibrational circular dichroism (VCD)] regions. In the presence of Mn2+, the protein-DNA interactions differ from those without the ions and cause prominent DNA compaction and formation of large intermolecular complexes. At the same time, the presence of HMGB1 and H1 also changed the mode of interaction of Mn2+ with DNA, which now takes place mostly in the major groove of DNA involving N7(G), whereas interactions between Mn2+ and DNA phosphate groups are weakened by histone molecules. Considerable interactions were also detected of Mn2+ ions with aspartic and glutamic amino acid residues of the proteins.  相似文献   

10.
Lipase from Pseudomonas cepacia was made soluble in 1,4‐dioxane by lyophilization of the enzyme from aqueous solutions containing methoxypoly(ethylene glycol) (PEG). The solubility of the enzyme–PEG complex depended both on protein concentration and PEG protein ratio. Intrinsic protein fluorescence and far‐ and near‐UV circular dichroism revealed that not only did the enzyme not unfold in the organic solvent, but rather became more compact. This was seen by the slight quenching of fluorescence intensity and by the enhancement of the near‐UV circular dichroism negative signals, which are indicative of stronger interactions of tryptophanyl and/or tyrosyl residues among themselves or with other parts of the enzyme molecule. The specific activity of the lipase–PEG complex in the organic solvent was at least 2 orders of magnitude higher than that of the enzyme powder. This can be attributed both to the maintenance of native conformation and to enzyme dissolution in the reaction medium which should minimize possible limitations to enzyme–substrate interactions. © 1999 John Wiley & Sons, Inc., Biotechnol Bioeng 64: 624–629, 1999.  相似文献   

11.
Nový J  Urbanová M 《Biopolymers》2007,85(4):349-358
The interactions of two different porphyrins, without axial ligands-5,10,15,20-tetrakis(1-methylpyridinium-4-yl)porphyrin-Cu(II) tetrachloride (Cu(II)TMPyP) and with bulky meso substituents-5,10,15,20-tetrakis(N,N,N-trimethylanilinium-4-yl)porphyrin tetrachloride (TMAP), with (dG-dC)10 and (dA-dT)10 were studied by combination of vibrational circular dichroism (VCD) and electronic circular dichroism (ECD) spectroscopy at different [oligonucleotide]/[porphyrin] ratios, where [oligonucleotide] and [porphyrin] are the concentrations of oligonucleotide per base-pair and porphyrin, respectively. The combination of VCD and ECD spectroscopy enables us to identify the types of interactions, and to specify the sites of interactions: The intercalative binding mode of Cu(II)TMPyP with (dG-dC)(10), which has been well described, was characterized by a new VCD "marker" and it was shown that the interaction of Cu(II)TMPyP with (dA-dT)10 via external binding to the phosphate backbone and major groove binding caused transition from the B to the non-B conformer. TMAP interacted with the major groove of (dG-dC)10, was semi-intercalated into (dA-dT)10, and caused significant variation in the structure of both oligonucleotides at the higher concentration of porphyrin. The spectroscopic techniques used in this study revealed that porphyrin binding with AT sequences caused substantial variation of the DNA structure. It was shown that VCD spectroscopy is an effective tool for the conformational studies of nucleic acid-porphyrin complexes in solution.  相似文献   

12.
In DNA, i‐motif (iM) folds occur under slightly acidic conditions when sequences rich in 2′‐deoxycytidine (dC) nucleotides adopt consecutive dC self base pairs. The pH stability of an iM is defined by the midpoint in the pH transition (pHT) between the folded and unfolded states. Two different experiments to determine pHT values via circular dichroism (CD) spectroscopy were performed on poly‐dC iMs of length 15, 19, or 23 nucleotides. These experiments demonstrate two points: (1) pHT values were dependent on the titration experiment performed, and (2) pH‐induced denaturing or annealing processes produced isothermal hysteresis in the pHT values. These results in tandem with model iMs with judicious mutations of dC to thymidine to favor particular folds found the hysteresis was maximal for the shorter poly‐dC iMs and those with an even number of base pairs, while the hysteresis was minimal for longer poly‐dC iMs and those with an odd number of base pairs. Experiments to follow the iM folding via thermal changes identified thermal hysteresis between the denaturing and annealing cycles. Similar trends were found to those observed in the CD experiments. The results demonstrate that the method of iM analysis can impact the pHT parameter measured, and hysteresis was observed in the pHT and Tm values.  相似文献   

13.
The continuously growing interest in the understanding of peptide folding led to the conformational investigation of methylamides of N‐acetyl‐amino acids as diamide models. Here we report the results of detailed conformational analysis on Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe diamides. These compounds were analyzed by experimental and computational methods, the conformational distributions obtained by Density Functional Theory (DFT) calculations for isolated and solvated diamide compounds are discussed. The conformational preference of proline‐containing diamide compounds as a function of the ambience was observed by a number of chiroptical spectroscopic techniques, such as vibrational circular dichroism (VCD), electronic circular dichroism (ECD), Raman optical activity (ROA) spectroscopy, and additionally by single crystal X‐ray diffraction analyses. Based on a comparison between Ac‐Pro‐NHMe and Ac‐β‐HPro‐NHMe, one can conclude that due to the greater conformational freedom of the β‐HPro derivative, Ac‐β‐HPro‐NHMe shows different behavior in solid‐ and solution‐phase, as well. Ac‐β‐HPro‐NHMe tends to form cis Ac‐β‐HPro amide conformation in water, dichloromethane, and acetonitrile in contrast to its α‐Pro analog. On the other hand, the crystal structure of the β‐HPro compound cannot be related to any of the conformers obtained in vacuum and solution while the X‐ray structure of Ac‐Pro‐NHMe was identified as tαL–, which is a trans Ac‐Pro amide containing conformer also predominant in polar solvents. Chirality 26:228–242, 2014. © 2014 Wiley Periodicals, Inc.  相似文献   

14.
The circular dichroism of Ac-(Ala)x-OMe and H-Lys-(Lys)x-OH with x = 1, 2, 3, and 4 has been measured in aqueous solutions. The oligomers with x = 4 show similar circular dichroism spectra in water when the lysyl amino groups are protonated, and they respond in similar fashion to heating and to sodium perchlorate. Both oligomers at 15°C exhibit a positive circular dichroism band at 217–218 nm, which is eliminated by the isothermal addition of 4 M sodium perchlorate or by heating. The positive circular dichroism of the lysine oligomer is also eliminated when the pH is elevated to deprotonate the amino groups. Positive circular dichroism is still observed for Ac-(Ala)4-OMe at elevated pH. Circular dichroism spectra have been estimated for poly(L -alanine) and poly(L -lysine) as statistical coils under the above conditions, based on the trends established with the oligomers. Poly(L -lysine) and poly(L -alanine) are predicted to exhibit similar circular dichroism behavior in aqueous solution so long as the lysyl amino groups are protonated. The circular dichroism of the statistical coil of poly(L -lysine), but not poly(L -alanine), is predicted to change when the pH is elevated sufficiently to deprotonate the lysyl amino groups. These results suggest that the unionized lysyl side chains participate in interactions that are not available to poly(L -alanine). Hydrophobic interactions may occur between the unionized lysyl side chains. Protonation of the lysyl amino groups is proposed to disrupt these interactions, causing poly(L -alanine) and protonated poly(L -lysine) to have similar circular dichroism properties.  相似文献   

15.
In this study, we performed an ultraviolet-visible (UV-Vis) and circular dichroism (CD) spectroscopic analysis of the binary and ternary supramolecular structures formed by self-assembling the following three water-soluble porphyrins with and without a chiral template: the negatively charged, meso-Tetra(4-sulfonatophenyl) porphine (H2TPPS4−); the positively charged meso-trans-(di(N-methyl-4-pyridyl)diphenyl) porphine (trans-DmPyDPP) and meso-cis-(di(N-methyl-4-pyridyl)diphenyl) porphine (cis-DmPyDPP). Polyglutamic acid (both L and D enantiomers) was selected as the chiral template due to its ability to change secondary structure with pH. The propensity for the porphyrins to show an induced CD in the presence of polyglutamic acid is demonstrated. The induced chirality of all supramolecular structures was found to depend on the pH of the solution, the chirality of the polymer, and the order of addition of the positively and negatively charged porphyrins (for ternary complexes). Of particular interest is that the interaction of H2TPPS4− with the chiral scaffold seems to undergo a dynamic rearrangement of the supramolecular structure as evident from the change in the CD spectrum over time. Moreover, experiments with ternary complexes suggest that the preferential interaction of trans-DmPyDPP with the random coil of the polymer shows promise as a sensor of protein secondary structure.  相似文献   

16.
Parallel PNA:PNA duplexes were synthesized and conjugated with meso‐tris(pyridyl)phenylporphyrin carboxylic acid at the N‐terminus. The introduction of one porphyrin unit was shown to affect slightly the stability of the PNA:PNA parallel duplex, whereas the presence of two porphyrin units at the same end resulted in a dramatic increase of the melting temperature, accompanied by hysteresis between melting and cooling curves. The circular dichroism (CD) profile of the Soret band and fluorescence quenching strongly support the occurrence of a face‐to‐face interaction between the two porphyrin units. Introduction of a L‐lysine residue at the C‐terminal of one strand of the parallel duplex induced a left‐handed helical structure in the PNA:PNA duplex if the latter contains only one or no porphyrin moiety. The left‐handed helicity was revealed by nucleobase CD profile at 240–280 nm and by the induced‐CD observed in the presence of the DiSC2(5) cyanine dye at ~500–550 nm. Surprisingly, the presence of two porphyrin units led to the disappearance of the nucleobase CD signal and the absence of CD exciton coupling within the Soret band region. In addition, a dramatic decrease of induced CD of DiSC2(5) was observed. These results are in agreement with a model where the porphyrin–porphyrin interactions cause partial loss of chirality of the PNA:PNA parallel duplex, forcing it to adopt a ladder‐like conformation. Chirality 27:864–874, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

17.
《Chirality》2017,29(11):653-662
The primary purpose of this work was to demonstrate the suitability of circular dichroism (CD) spectroscopy in stereochemical studies of α,β ‐unsaturated oximes, with particular emphasis on determination of E and Z geometry of the oxime double bond. As models for this study, O‐phenyl and O‐triphenylmethyl (trityl) oximes of 4‐hydroxy‐2‐methylcyclopent‐2‐en‐1‐one were selected. These model compounds differ in both absolute configuration at C4 carbon atom and E Z configuration of the oxime double bond. The basic dichroic technique applied was electronic circular dichroism (ECD) assisted by quantum‐chemical calculations and vibrational circular dichroism (VCD) for selected cases. Such an approach enabled effective implementation of both goals. Thus, we were able to associate the signs of Cotton effects in the range of 190–240 nm with the absolute configuration at C4 and within 240–300 nm with the E ‐ or Z ‐geometry of the oxime double bond. Within this work, optical activity of the protecting trityl group was also studied towards formation of the propeller‐shaped conformations by using the same combined CD/DFT methodology. As shown, the helical structure of the trityl group has a considerable influence on the ECD spectra. However, the MPM and PMP conformers of the trityl group are in fact almost equally populated in the conformational equilibrium, making it impossible to distinguish them. On the other hand, rotamers of the hydroxyl group at C4 show a decisive impact on the VCD spectra in both phenoxy and trityl oximes.  相似文献   

18.
The electronic circular dichroism (ECD) spectrum of the recently synthesized [16]helicene and a derivative comprising two triisopropylsilyloxy protection groups was computed by means of the very efficient simplified time‐dependent density functional theory (sTD‐DFT) approach. Different from many previous ECD studies of helicenes, nonequilibrium structure effects were accounted for by computing ECD spectra on "snapshots" obtained from a molecular dynamics (MD) simulation including solvent molecules. The trajectories are based on a molecule specific classical potential as obtained from the recently developed quantum chemically derived force field (QMDFF) scheme. The reduced computational cost in the MD simulation due to the use of the QMDFF (compared to ab‐initio MD) as well as the sTD‐DFT approach make realistic spectral simulations feasible for these compounds that comprise more than 100 atoms. While the ECD spectra of [16]helicene and its derivative computed vertically on the respective gas phase, equilibrium geometries show noticeable differences, these are “washed” out when nonequilibrium structures are taken into account. The computed spectra with two recommended density functionals (ωB97X and BHLYP) and extended basis sets compare very well with the experimental one. In addition we provide an estimate for the missing absolute intensities of the latter. The approach presented here could also be used in future studies to capture nonequilibrium effects, but also to systematically average ECD spectra over different conformations in more flexible molecules. Chirality 28:365–369, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

19.
Marcin Górecki 《Chirality》2015,27(7):441-448
Recent advances in equipment enabled the collection of solid‐state electronic circular dichroism (ECD) spectra using the commercially available integrating sphere attachment for a regular ECD spectrometer. This accessory was designed to reduce negative factors occurring in solid‐state ECD measurements, and is, thereby, very useful for recording diffuse transmittance CD (DTCD) spectra using the pellet technique. In the present article, the operating principle of the integrating sphere and utility of the DTCD method in recording solid‐state ECD spectra is demonstrated. Based on illustrative examples, i.e., 10‐camphorsulfonic acid ammonium, cholest‐4‐en‐3‐one, (3R,4R,5S)‐oseltamivir, and (S)‐linezolid, ECD solid‐state measurements were performed by means of both transmission and diffusion methods and later compared. Selection of these compounds as models for comparative studies was made in view of their different chromophoric systems and the profound importance in the pharmaceutical industry. During the course of this work the benefits and limitations of the use of integrating sphere are presented. The final conclusion is that more relevant solid‐state spectra can be obtained by means of the DTCD method. Chirality 27:441–448, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

20.
The chiral separation of enantiomeric couples of three potential A3 adenosine receptor antagonists: (R/S)‐N‐(6‐(1‐phenylethoxy)‐2‐(propylthio)pyrimidin‐4‐yl)acetamide ( 1 ), (R/S)‐N‐(2‐(1‐phenylethylthio)‐6‐propoxypyrimidin‐4‐yl)acetamide ( 2 ), and (R/S)‐N‐(2‐(benzylthio)‐6‐sec‐butoxypyrimidin‐4‐yl)acetamide ( 3 ) was achieved by high‐performance liquid chromatography (HPLC). Three types of chiroptical spectroscopies, namely, optical rotatory dispersion (ORD), electronic circular dichroism (ECD), and vibrational circular dichroism (VCD), were applied to enantiomeric compounds. Through comparison with Density Functional Theory (DFT) calculations, encompassing extensive conformational analysis, full assignment of the absolute configuration (AC) for the three sets of compounds was obtained. Chirality 28:434–440, 2016. © 2016 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