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.
Because of the generally different interaction of enantiomers with biological systems, there has been an ever increasing demand for artificial highly enantioselective systems that can facilitate separation processes involved in the research and development of enantiomerically pure drugs. Such systems may be discovered by large‐scale screening of compound libraries which warrants rapid and cost‐efficient screening methods. Here, we demonstrate enantioselectivity determination for systems of cinchona alkaloid carbamates and N‐blocked amino acids using HPLC‐MS and the recently developed dynamic titration technique (Fry?ák P, Schug KA. Anal Chem 2008;80:1385‐1393). A mixture of nine N‐blocked amino acids (either D or L enantiomers) was separated on a reversed‐phase column with cinchona alkaloid carbamates added postcolumn. Dissociation constants of the observed noncovalent complexes were determined from the HPLC‐MS data. Enantioselectivity was then calculated from the dissociation constants, pointing out the best performing systems. For these systems, apparent dissociation constants were measured for the whole range of enantiomeric composition and were shown to obey a proposed theoretical model. Chirality, 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

2.
The temperature-dependence of a large number of NMR parameters describing hydrogen bond properties in the protein ubiquitin was followed over a range from 5 to 65 degrees C. The parameters comprise hydrogen bond (H-bond) scalar couplings, h3JNC', chemical shifts, amide proton exchange rates, 15N relaxation parameters as well as covalent 1JNC' and 1JNH couplings. A global weakening of the h3JNC' coupling with increasing temperature is accompanied by a global upfield shift of the amide protons and a decrease of the sequential 1JNC' couplings. If interpreted as a linear increase of the N...O distance, the change in h3JNC' corresponds to an average linear thermal expansion coefficient for the NH-->O hydrogen bonds of 1.7 x 10(-4)/K, which is in good agreement with overall volume expansion coefficients observed for proteins. A residue-specific analysis reveals that not all hydrogen bonds are affected to the same extent by the thermal expansion. The end of beta-sheet beta1/beta5 at hydrogen bond E64-->Q2 appears as the most thermolabile, whereas the adjacent hydrogen bond I3-->L15 connecting beta-strands beta1 and beta2 is even stabilized slightly at higher temperatures. Additional evidence for the stabilization of the beta1/beta2 beta-hairpin at higher temperatures is found in reduced hydrogen exchange rates for strand end residue V17. This reduction corresponds to a stabilizing change in free energy of 9.7 kJ/mol for the beta1/beta2 hairpin. The result can be linked to the finding that the beta1/beta2 hairpin behaves as an autonomously folding unit in the A-state of ubiquitin under changed solvent conditions. For several amide groups the temperature-dependencies of the amide exchange rates and H-bond scalar couplings are uncorrelated. Therefore, amide exchange rates are not a sole function of the hydrogen bond "strength" as given by the electronic overlap of donors and acceptors, but are clearly dependent on other blocking mechanisms.  相似文献   

3.
A technique based exclusively on chiral reversed-phase liquid chromatography has been shown to greatly facilitate studies of enantioselectivity in lipase-catalyzed hydrolysis of chiral organic esters. Only two sets of experimental data are needed to calculate the enantioselectivity (E) of a kinetically controlled enantiomer-differentiating reaction of this kind, viz. the enantiomeric excess of the product (eep) or substrate (ees), and the degree of substrate conversion (c). The product enantiomers are well separated on a BSA-based column, giving eep directly. In addition, separation of the (unresolved) ester substrate from the enantiomeric products gives c by integration. Via an optimization of the mobile phase used in the chiral chromatographic system, both these parameters can often be determined in a single run. Highly precise and detailed kinetic studies of the enzymatic reaction can thus be performed. In this way, E-values have been determined for a series of 2-chloroethyl 2-arylpropanoates hydrolyzed in the presence of a Candida cylindracea lipase at pH 6.0 and 7.1. Effects on the E-values from a partial purification and further processing of the lipase have also been studied.  相似文献   

4.
The mitochondrial steroid hydroxylase system of vertebrates utilizes adrenodoxin (Adx), a small iron–sulfur cluster protein of about 14 kDa as an electron carrier between a reductase and cytochrome P450. Although the crystal structure of this protein has been elucidated, the solution structure of Adx was discussed contrary in the literature [I.A. Pikuleva, K. Tesh, M.R. Waterman, Y. Kim, The tertiary structure of full-length bovine adrenodoxin suggests functional dimers, Arch. Biochem. Biophys. 373 (2000) 44–55; D. Beilke, R. Weiss, F. Löhr, P. Pristovsek, F. Hannemann, R. Bernhardt, H. Rüterjans, A new electron mechanism in mitochondrial steroid hydroxylase systems based on structural changes upon the reduction of adrenodoxin, Biochemistry 41 (2002) 7969–7978]. Therefore, it was necessary to study the self-association of this protein by using analytical ultracentrifugation over a larger concentration range. As could be demonstrated in sedimentation velocity experiments, as well as sedimentation equilibrium runs with explicit consideration of thermodynamic non-ideality, the full-length protein (residues 1–128) in the oxidized state resulted in a monomer–dimer equilibrium (Ka ~ 3 × 102 M− 1). For truncated Adx (1–108), as well as the reduced Adx, the association behavior was strongly reduced. The consequences of this behavior are discussed with respect to the physiological meaning for the Adx system.  相似文献   

5.
Chemical shift mapping is becoming a popular method for studying protein-protein interactions in solution. The technique is used to identify putative sites of interaction on a protein surface by detecting chemical shift perturbations in simple (1H, 15N)-HSQC NMR spectra of a uniformly labeled protein as a function of added (unlabeled) target protein. The high concentrations required for these experiments raise questions concerning the possibility for non-specific interactions being detected, thereby compromising the information obtained. We demonstrate here that the simple chemical shift mapping approach faithfully reproduces the known functional specificities among pairs of closely related proteins from the phosphoenolpyruvate:sugar phosphotransferase systems of Escherichia coli and Bacillus subtilis.  相似文献   

6.
The direct HPLC separation of eight inherently chiral atropisomeric calix[4]arenes has been achieved using Chiralcel OD phase. A rationale is given for the variation of the enantioselectivity as a function of the O-alkyl or O-aryl groups. In closely related structures hydrogen bond formation between the free hydroxyl of the analyte and the chiral phase plays an important role in the chiral recognition process. © 1993 Wiley-Liss, Inc.  相似文献   

7.
8.
Im SH  Ryoo JJ  Lee KP  Choi SH  Jeong YH  Jung YS  Hyun MH 《Chirality》2002,14(4):329-333
Recently, it was reported that the chiral recognition ability of (R)-N-3,5-dinitrobenzoyl phenylglycinol derivative was examined as a new HPLC chiral stationary phase (CSP 1) for the resolution of racemic N-acylnaphthylalkylamines. However, the mechanism of chiral discrimination on the CSP remained elusive until now. In this study, a spectroscopic investigation of the chiral discrimination mechanism of CSP 1 was undertaken using mixtures of (R)-N-3,5-dinitrobenzoyl phenylglycinol-derived chiral selector (2) and each of the enantiomers of N-acylnaphthylalkylamines (3) by NMR study. First, the differences in free energy changes (DeltaDeltaG) upon diastereomeric complexation in solution between the complex of each isomer with chiral selector 2 by NMR titration were calculated. The values were then compared with those estimated by chiral HPLC. The chemical shift changes of each proton on the chiral selector and analytes were also checked and it was found that the chemical shift changes decreased continuously as the acyl group on analytes increased in length. This observation was consistent with the HPLC data. From these experimental results, the interaction mechanism of chiral discrimination between the chiral selector and the analytes is more precisely explained.  相似文献   

9.
The conformation of an elastin-mimetic recombinant protein, [(VPGVG)4(VPGKG)]39, is investigated using solid-state NMR spectroscopy. The protein is extensively labeled with 13C and 15N, and two-dimensional 13C-13C and 15N-13C correlation experiments were carried out to resolve and assign the isotropic chemical shifts of the various sites. The Pro 15N, 13Calpha, and 13Cbeta isotropic shifts, and the Gly-3 Calpha isotropic and anisotropic chemical shifts support the predominance of type-II beta-turn structure at the Pro-Gly pair but reject a type-I beta-turn. The Val-1 preceding Pro adopts mostly beta-sheet torsion angles, while the Val-4 chemical shifts are intermediate between those of helix and sheet. The protein exhibits a significant conformational distribution, shown by the broad line widths of the 15N and 13C spectra. The average chemical shifts of the solid protein are similar to the values in solution, suggesting that the low-hydration polypeptide maintains the same conformation as in solution. The ability to measure these conformational restraints by solid-state NMR opens the possibility of determining the detailed structure of this class of fibrous proteins through torsion angles and distances.  相似文献   

10.
The enantiomeric separation of several racemic aryloxyaminopropan-2-ol derivatives related to propranolol on normal and reversed phase of cellulose tris (3,5-dimethylphenylcarbamate) chiral stationary phases known as Chiralcel OD and Chiralcel OD-R were studied. It was observed that the chiral separation depends on the substitution pattern of the aryl group, i.e., 1-naphthyl, 2-naphthyl, and phenyl group and polarity on the basic nitrogen in the side chain. In both normal and reversed phase modes the (+)-R-enantiomer eluted first in all of the analogs resolved. It can be concluded that: (1) substituents on the side chain did affect the interaction of the enantiomers with the polar carbamate moiety in the CSP; and (2) the dipole-dipole stacking between the π-donor 3,5-dimethylphenyl carbamate group pending from the glucose rings of the CSP and π-acceptor aryl group of the analyte is crucial for the efficient chiral discrimination. The chiral recognition mechanism(s) between these analogs and the chiral stationary phases are proposed. © 1996 Wiley-Liss, Inc.  相似文献   

11.
Identifying potential ligand binding sites on a protein surface is an important first step for targeted structure-based drug discovery. While performing control experiments with Escherichia coli peptide deformylase (PDF), we noted that the organic solvents used to solubilize some ligands perturbed many of the same resonances in PDF as the small molecule inhibitors. To further explore this observation, we recorded (15)N HSQC spectra of E. coli peptide deformylase (PDF) in the presence of trace quantities of several simple organic solvents (acetone, DMSO, ethanol, isopropanol) and identified their sites of interaction from local perturbation of amide chemical shifts. Analysis of the protein surface structure revealed that the ligand-induced shift perturbations map to the active site and one additional surface pocket. The correlation between sites of solvent and inhibitor binding highlights the utility of organic solvents to rapidly and effectively validate and characterize binding sites on proteins prior to designing a drug discovery screen. Further, the solvent-induced perturbations have implications for the use of organic solvents to dissolve candidate ligands in NMR-based screens.  相似文献   

12.
Synthetic soluble (—)-dopa melanin was prepared in deuteriated buffer, pH 8, by autooxidation of the precursor. At 6 mM of the precursor, the incorporation was over 90%. The changes in the line width measurements of N-CH3 protons of enantiomers of ephedrine in the soluble melanin were quantified by NMR spectroscopy. The dissociation constants of (—)-1R,2S-ephedrine, (+)-1S,2R-ephedrine, (—)-1R,2R-ψ-ephedrine, and (+)-1S,2S-ψ-ephedrine were 11.7, 4.20, 3.60, and 4.80 mM, respectively. Since the concentration of (—)-dopa was known and since the conversion of (—)-dopa to indole units of melanin was considered as 1:1, the stoichiometry of the interaction between the drug and the indole unit was calculated. Based on the dissociation constants of the enantiomers, it appears that up to four molecules of (—)-ephedrine can interact with one indole unit of the melanin, while such a ratio for other isomers appear to be 2:1. The preference by indole units of melanin is stereoselective. © 1992 Wiley-Liss, Inc.  相似文献   

13.
Ding H  Yang Y  Zhang J  Wu J  Liu H  Shi Y 《Proteins》2005,61(4):1050-1058
The interaction between small ubiquitin-related modifier SUMO and its conjugating-enzyme Ubc9 (E2) is an essential step in SUMO conjugation cascade. However, an experimental structure of such a transient complex is still unavailable. Here, a structural model of SUMO-3-Ubc9 complex was obtained with HADDOCK, combining NMR chemical shift mapping information. Docking calculations were performed using SUMO-3 and Ubc9 structures as input. The resulting complex reveals that the complementary surface electrostatic potentials contribute dominantly to the specific interaction. At the interface, similar numbers of oppositely-charged conserved residues are identified on the respective binding partners. Hydrogen bonds are formed in the vicinity of the interface to stabilize the complex. Comparison of the structure of SUMO-3-Ubc9 complex generated by HADDOCK and the experimental structures in free form indicates that SUMO-3 and Ubc9 maintain their respective fold as a whole after docking. However, the N-terminal helix alpha1 and its subsequent L1 loop of Ubc9 experience sizeable changes upon complex formation. They cooperatively move towards the hydrophilic side of the beta-sheet of SUMO-3. Our observations are consistent with the data from previous Ubc9 mutational analysis and conformational flexibility studies. Together, we have proposed that the SUMO-3-Ubc9 interaction is strongly electrostatically driven and the N terminus of Ubc9 shifts to SUMO-3 to facilitate the interaction. The NMR-based structural model, which provides considerable insights into the molecular basis of the specific SUMO-E2 recognition and interaction, implicates the general interaction mode between SUMO-3 and Ubc9 homologues from yeast to humans.  相似文献   

14.
Paramagnetic relaxation enhancement provides a tool for studying the dynamics as well as the structure of macromolecular complexes. The application of side-chain coupled spin-labels is limited by the mobility of the free radical. The cyclic, rigid amino acid spin-label TOAC (2,2,6,6-Tetramethylpiperidine-1-oxyl-4-amino-4-carboxylic acid), which can be incorporated straightforwardly by peptide synthesis, provides an attractive alternative. In this study, TOAC was incorporated into a peptide derived from focal adhesion kinase (FAK), and the interaction of the peptide with the Src homology 3 (SH3) domain of Src kinase was studied, using paramagnetic NMR. Placing TOAC within the binding motif of the peptide has a considerable effect on the peptide-protein binding, lowering the affinity substantially. When the TOAC is positioned just outside the binding motif, the binding constant remains nearly unaffected. Although the SH3 domain binds weakly and transiently to proline-rich peptides from FAK, the interaction is not very dynamic and the relative position of the spin-label to the protein is well-defined. It is concluded that TOAC can be used to generate reliable paramagnetic NMR restraints.  相似文献   

15.
Qiu L  Wang Q  Lin L  Liu X  Jiang X  Zhao Q  Hu G  Wang R 《Chirality》2009,21(2):316-323
A new catalytic system, generated from the readily available and inexpensive beta-sulfonamide alcohol L*, Ti(O(i)Pr)(4), Et(2)Zn, and tertiary amine base (R(3)N), effectively catalyzes the enantioselective addition of various terminal alkynes including some quite challenging alkynes to aldehydes in good yields and excellent enantioselectivities. Up to 96% yield and >99% enantioselectivity were achieved with the use of N,N-diisoproylethylamine (DIPEA) as an additive in this asymmetric addition.  相似文献   

16.
《Chirality》2017,29(11):747-758
Gibbs energies of complex formation between enantiomers of bicyclic terpenoid, fenchone, and naturally occurring cyclodextrins, βCD and γCD, were determined by means of 13C and 1H nuclear magnetic resonance (NMR) titration data. These results were compared with the corresponding data obtained previously for the diastereomeric fenchone/αCD complexes. The size of the inner cavity of host molecules significantly influences stoichiometry, association constants, and enantiomeric differentiation of the studied complexes. These complementary data allow us to discuss qualitatively the influence of the host size on the guest–host interactions. A method of the simultaneous use of titration data collected for several resonances of different isotopes in the determination of association constants was worked out and thoroughly analyzed. Comparison of the results of global data analyses with weighted means of individual ones revealed that both these approaches are equally trustworthy.  相似文献   

17.
The inclusion complexation behavior of chiral members of cinchona alkaloid with beta- and gamma-cyclodextrins (1 and 2) and 6,6(')-trimethylenediseleno-bridged bis(beta-cyclodextrin) (3) was assessed by means of fluorescence and 2D-NMR spectroscopy. The spectrofluorometric titrations have been performed in aqueous buffer solution (pH 7.20) at 25.0 degrees C to determine the stability constants of the inclusion complexation of 1-3 with guest molecules (i.e., cinchonine, cinchonidine, quinine, and quinidine) in order to quantitatively investigate the molecular selective binding ability. The stability constants of the resulting complexes of 2 with guest molecules are larger than that of 1. As a result of cooperative binding, the stability constants of inclusion complexation of dimeric beta-cyclodextrin 3 with cinchonidine and cinchonine are higher than that of parent 1 by factor of 4.5 and 2.4, respectively. These results are discussed from the viewpoint of the size-fit and geometric complementary relationship between the host and guest.  相似文献   

18.
We introduce a solid-state NMR technique for selective detection of a residue pair in multiply labeled proteins to obtain site-specific structural constraints. The method exploits the frequency-offset dependence of cross polarization to achieve 13COi 15Ni 13Ci transfer between two residues. A 13C, 15N-labeled elastin mimetic protein (VPGVG)n is used to demonstrate the method. The technique selected the Gly3 C signal while suppressing the Gly5 C signal, and allowed the measurement of the Gly3 C chemical shift anisotropy to derive information on the protein conformation. This residue-pair selection technique should simplify the study of protein structure at specific residues.  相似文献   

19.
Cisproline(i - 1)-aromatic(i) interactions have been detected in several short peptides in aqueous solution by analysis of anomalous chemical shifts measured by 1H-NMR spectroscopy. This formation of local structure is of importance for protein folding and binding properties. To obtain an atomic-detail characterisation of the cisproline(i - 1)-aromatic(i) interaction in terms of structure, energetics and dynamics, we studied the minimal peptide unit, blocked Ala-cisPro-Tyr, using computational and experimental techniques. Structural database analyses and a systematic search revealed two groups of conformations displaying a cisproline(i - 1)-aromatic(i) interaction. These conformations were taken as seeds for molecular dynamics simulations in explicit solvent at 278 K. During a total of 33.6 ns of simulation, all the 'folded' conformations and some 'unfolded' states were sampled. 1H- and 13C-chemical shifts and 3J-coupling constants were measured for the Ala-Pro-Tyr peptide. Excellent agreement was found between all the measured and computed NMR properties, showing the good quality of the force field. We find that under the experimental and simulation conditions, the Ala-cisPro-Tyr peptide is folded 90% of the time and displays two types of folded conformation which we denote 'a' and 'b'. The type a conformations are twice as populated as the type b conformations. The former have the tyrosine ring interacting with the alanine alpha proton and are enthalpically stabilised. The latter have the aromatic ring interacting with the proline side chain and are entropically stabilised. The combined and complementary use of computational and experimental techniques permitted derivation of a detailed scenario of the 'folding' of this peptide.  相似文献   

20.
Capillary electrophoresis (CE) allows the observation of the opposite affinities of the enantiomers of (±)‐verapamil [2‐isopropyl‐2,8‐bis(3,4‐dimethoxyphenyl)‐6‐methyl‐6‐azaoctannitrile, VP] toward β‐cyclodextrin (β‐CD) and heptakis(2,3,6‐tri‐O‐methyl)‐β‐CD (TM‐β‐CD). In addition, in the presence of β‐CD in the background electrolyte, longer migration times and lower separation factors were observed compared to TM‐β‐CD. The binding constants of (+)‐ and (−)‐VP with β‐CD and TM‐β‐CD determined using 13C NMR spectroscopy explain the results observed in CE. Electrospray ionization mass spectrometry (ESI‐MS) was used as an alternative technique for the characterization of VP‐CD complexes. Chirality 11:635–644, 1999. © 1999 Wiley‐Liss, 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