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 46 毫秒
1.
Herein we report an efficient one pot synthesis of new chiral 4,5‐dihydro‐4‐arylspiro[1,3,4‐thiadiazole]‐5,2′‐camphane‐2‐carboxylic acid ethyl esters 5–7 and 4,5‐dihydro‐3‐arylspiro[1,4,2‐oxathiazole]‐5,2′‐camphane 11–13 , using 1,3‐dipolar cycloaddition of nitrilimines 2–4 and nitrile oxides 8–10 to (1R)‐thiocamphor 1 respectively. The structure of the newly prepared 1,3,4‐thiadiazoles 5–7 (obtained as pure diastereoisomers) were fully established via spectroscopic analysis and X‐ray structural analysis which proved the absolute configuration of the C5 spiranic carbon to be (R). NMR spectral analysis were also very useful to show the new 1,4,2‐oxathiazoles 11–13 are mixtures of two (5R)/(5S) diastereoisomers with the ratio 6:4,7:3 and 6:4 respectively.  相似文献   

2.
4-(Phenylamino)-pyrrolo[2,1-f][1,2,4]triazines have been discovered as inhibitors of p38α. Experimental assays have proven that the configuration of α-Me-benzyl connected with amide at C6 is essential for the binding affinity. The S-configured inhibitor (11j) displays 80 times more potency than the R-configured one (11k). Here we investigated the mechanism how different configurations influence the binding affinity using molecular dynamics simulations, free energy calculations and free energy decomposition analysis. We found that the van der Waals interactions play the most important role in differentiating the activities between 11j and 11k with p38α. The difference of the van der Waals interactions is primarily determined by two residues, LEU108 and LEU167. Consequently stabilization of pyrrolo[2,1-f][1,2,4]triazine ring is important for the activities of inhibitors. Meanwhile we observed that the different configuration of the α-Me-benzyl group leads to the difference of binding between 11j and 11k. In conclusion, our work shows that it is feasible to analyze the chirality effect of inhibitors with different configurations by molecular dynamics simulations and free energy calculations, and provides useful information for drug design.  相似文献   

3.
Analytical results of a series of poly(N-isopropylacrylamide) (PNiPA) with different tacticities using infrared spectroscopy are presented for studying the influence of the solvation and molecular interactions between the polymeric chains. Infrared spectra of solid matter samples of the compounds exhibit a systematic band intensity change for three band components at 1,680, 1,659, and 1,628 cm−1 involved in the amide I band. The three components correspond to the free, half, and full hydrogen bondings of the secondary amide group, which reflect the molecular configuration depending on the tacticity. When cast films of the compounds prepared on a solid surface are analyzed by infrared transmission spectrometry, another factor of the solvent used for the film preparation is found to be another factor which plays an important role in determining the molecular architecture in the films. This molecular imprint mechanism after the solvation is confirmed by measuring infrared multiple-angle incidence resolution spectra of annealed films. The molecular interactions in the polymeric samples have been revealed by the use of infrared spectroscopy and the tacticity-controlled samples.  相似文献   

4.
A general procedure to determine the absolute configuration of cyclic secondary amines with Mosher's NMR method is demonstrated, with assignment of absolute configuration of isoanabasine as an example. Each Mosher amide can adopt two stable conformations (named rotamers) caused by hindered rotation around amide C--N bond. Via a three-step structural analysis of four rotamers, the absolute configuration of (-)-isoanabasine is deduced to be (R) on the basis of Newman projections, which makes it easy to understand and clarify the application of Mosher's method to cyclic secondary amines. Furthermore, it was observed that there was an unexpected ratio of rotamers of Mosher amide derived from (R)-isoanabasine and (R)-Mosher acid. This phenomenon implied that it is necessary to distinguish the predominant rotamer from the minor one prior to determining the absolute configuration while using this technique.  相似文献   

5.
Hapalosin was initially synthesized by macrolactonization, and a second synthesis was achieved by cycloamidation. In both syntheses, three of the five stereocenters in hapalosin were established by two Brown allylboration reactions. The synthesis of the non-N-Me analog of hapalosin involved chelation-controlled reduction of a gamma-amino-beta-keto ester and cycloamidation. In CDCl(3) at 25 degrees C, synthetic hapalosin exists as a 2.3:1 mixture of conformers, while its non-N-Me analog exists only as a single conformer. (1)H,(1)H-NOESY and computation reveal that the configuration of the amide bond is responsible for the conformations of the two compounds. The major conformer of hapalosin is found to be an s-cis amide, the minor conformer an s-trans amide, and the non-N-Me analog an s-trans amide. Applying distance constraints to protons that exhibit NOESY correlations, computation shows that the major conformer of hapalosin and the non-N-Me analog have very different conformations. By contrast, the minor conformer of hapalosin and the non-N-Me analog have very similar conformations.  相似文献   

6.
Because proteins adopt unique structures, chemically identical nuclei in proteins exhibit different chemical shifts. Amide 15N chemical shifts have been shown to vary over 20 ppm. The cause of these chemical shift inequivalencies is the different intra‐ and intermolecular interactions that individual nuclei experience at different locations in the protein structure. These chemical shift inequivalencies can be described as structural shifts, the difference between the actual chemical shift and the random coil chemical shift. As a first step toward the prediction of these amide 15N structural shifts, calculations have been carried out on acetyl‐glycine‐methyl amide to examine how a neighboring peptide group influences the amide 15N structural shifts. The ϕ,ψ dihedral angle space is completely surveyed, while all other geometrical variables are held fixed, to isolate the effect of the backbone conformation. Similar calculations for a limited number of conformations of acetyl‐glycine‐glycine‐methyl amide were carried out, where the effects of the two terminal peptide groups on the central amide 15N structural shift are examined. It is shown that the effect of the two adjacent groups can be accurately modeled by combining their individual effects additively. This provides a quite simple method to predict the backbone influence on amide 15N structural shifts in proteins. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 366–372, 2001  相似文献   

7.
《Liquid crystals》1999,26(3):389-396
Two series of calamitic liquid crystals containing a benzothiazole ring within the central core and two different linkage groups (amide and azo) have been prepared and their liquid crystalline properties studied and compared with those of the analogous series of imines. The influence of the linkage group within the central core has been proven to determine the variety of mesomorphism displayed by the compounds. The compounds with imine and azo linkages behave in a similar way and exhibit typical nematic and smectic C mesophases. Compounds incorporating an amide linkage show a poorer mesomorphism and mainly present a smectic C mesophase.  相似文献   

8.
王利敏  程森祥  陈彤  常俊标 《化学学报》2012,70(10):1201-1206
在(S)-氨基丙醇及(R)-氨基丙醇手性臂的作用下, (M/P)-4,4'-二甲氧基-5,6,5',6'-二次甲二氧基-2-甲酸酯-2'-甲酰氨联苯经构型转化, 制备了光学纯轴手性联苯化合物(P, S)-3a 和(M, R)-3a. 测定了(P, S)-3a 的晶体结构及CD 光谱. 结果表明, 化合物(P, S)-3a 晶体属单斜晶系, P2(1)空间群, 晶胞参数为a=12.122(2) Å, b=8.9911(18) Å, c=12.779(3) Å, β=112.38(3)°, 在晶体中存在两组分子间氢键相互作用, 一组氢键由羟基氢与另一分子酰胺基团的羰基氧组成O-H…O,另一组由酰胺基团的NH 与另一分子酰胺基团的羰基氧构成N-H…O, 每一分子通过四个氢键与另外两个分子相连,构成棒状结构. 由CD 光谱确定了(M, R)-3b 的立体构型. 此外, 由(P, S)-3a 合成了轴手性化合物(P)-2,2'-二羟甲基-4,4'-二甲氧基-5,6,5',6'-二亚甲二氧基-联苯(6a).  相似文献   

9.
A practical synthesis of new chiral aminophosphine ligands based on the camphane scaffold bearing alkoxy groups was accomplished. The application of these ligands in the Pd-catalyzed allylic alkylation of (E)-1,3-diphenyl-2-propenyl acetate proceeded with excellent conversions and ee’s of up to 91%. Variation of the alkoxy substituents did not substantially influence the catalytic performance.  相似文献   

10.
α,β-Unsaturated hydroxamates derived from the ‘chiral Weinreb amide’ auxiliary (S)-N-1-(1′-naphthyl)ethyl-O-tert-butylhydroxylamine consistently adopt a defined conformation and undergo highly diastereoselective conjugate addition reactions with lithium amide reagents. The configuration of the N-1-(1′-naphthyl)ethyl group dictates the position of the O-tert-butyl group and also the configuration adopted by the pyramidal nitrogen atom via a ‘chiral relay’ effect. Conjugate addition of lithium amide reagents to these substrates proceeds on the face opposite to both the O-tert-butyl group and nitrogen lone-pair with high levels of diastereoselectivity.  相似文献   

11.
Configurational assignments in a series of N-alkyl-N-benzylbenzamides using a variety of NMR techniques lead to consistent assignments in amides with two isomers but are inconclusive for amides with single isomers. Single crystal X-ray structure determination shows the configuration of o-bromo-N-benzyl-N-t-butyl-benzamide to have the t-Bu group syn to the carbonyl oxygen. Surprisingly, the major isomer of o-chloro-N-benzyl-N-i-propylbenzamide, appears to have the opposite configuration, i.e., the i-Pr group is anti to the carbonyl oxygen. The structural details of o-bromo-N-benzyl-N-t-butylbenzamide reveal slight deviation from planarity for the amide group and chirality of the benzoyl group.  相似文献   

12.
An enantiomerically pure (1-trimethylsilyl)ethyl group, constructed by a (-)-sparteine-directed enantioselective quench of a laterally lithiated tertiary aromatic amide, exerts powerful thermodynamic control over the conformation of the adjacent tertiary amide substituent. Ortholithiation and functionalization of the amide in the 6-position allows the single amide conformer to be trapped as an enantiomerically and diastereoisomerically pure amide atropisomer. Protodesilylation of the amide gives functionalized atropisomeric amides with a stereogenic axis of single absolute configuration, whose barriers to racemization have been determined by polarimetry. Enantiomerically pure amides bearing phosphine substituents are effective ligands in a Pd-catalyzed allylic substitution reaction-the first use of a nonbiaryl atropisomer as a chiral ligand-and give products with 90% ee. The rate of racemization of the phosphine-substituted amide is powerfully influenced by the presence of palladium.  相似文献   

13.
The molecular structure and IR spectra of urea, H2NCONH2, in gas phase and in acetonitrile solution, as well as of the two complexes [MgU4Cl2] and [MgU6]Cl2 have been observed. The influence of environmental changes to geometry and spectra are shown. Various basis sets have been employed to safeguard the validity of the reported findings, using polarization functions for all calculations to get the correct pyramidal amide configuration. The erroneous low energy of the C2v symmetry group, after the addition of the ZPVE correction, is discussed. For the solvated urea molecule a reduction of the energy barrier, compared to the gas phase urea, between the two minimum configurations, C2 and Cs, and the planar geometry, is observed. The lowest energy minimum in acetonitrile is found to be the C2 symmetry group, while for the two complexes, the local symmetry of urea is Cs or C2 depending on the complex, or even on the coordination position of urea in the complex. The wagging motion of the amide group is also discussed in all the studied urea species. The computed geometries and most of the spectroscopic results are in good agreement with the available experimental data. Received: 18 July 2000 / Accepted: 31 July 2000 / Published online: 2 November 2000  相似文献   

14.
One advantage of detecting amide H/2H exchange by mass spectrometry instead of NMR is that the more rapidly exchanging surface amides are still detectable. In this study, we present quench-flow amide H/2H exchange experiments to probe how rapidly the surfaces of two different proteins exchange. We compared the amide H/2H exchange behavior of thrombin, a globular protein, and IkappaBalpha, a nonglobular protein, to explore any differences in the determinants of amide H/2H exchange rates for each class of protein. The rates of exchange of only a few of the surface amides were as rapid as the "intrinsic" exchange rates measured for amides in unstructured peptides. Most of the surface amides exchanged at a slower rate, despite the fact that they were not seen to be hydrogen bonded to another protein group in the crystal structure. To elucidate the influence of the surface environment on amide H/2H exchange, we compared exchange data with the number of amides participating in hydrogen bonds with other protein groups and with the solvent accessible surface area. The best correlation with amide H/2H exchange was found with the total solvent accessible surface area, including side chains. In the case of the globular protein, the correlation was modest, whereas it was well correlated for the nonglobular protein. The nonglobular protein also showed a correlation between amide exchange and hydrogen bonding. These data suggest that other factors, such as complex dynamic behavior and surface burial, may alter the expected exchange rates in globular proteins more than in nonglobular proteins where all of the residues are near the surface.  相似文献   

15.
The influence of a hydrogen bond donor and acceptor in the hydrophobic part of an amphiphile on the monolayer stability at the air/water interface is investigated. For that purpose, the amide group is integrated into the alkyl chain. Eight methyl octadecanoates have been synthesized with the amide group in two orientations and in different positions of the alkyl chain, namely, CH3O2C(CH2)m NHCO(CH2)n CH3 (n + m = 14): 1 (m = 1), 3 (m = 2), 5 (m = 3), 7 (m = 14); and CH3O2C(CH2)m CONH(CH2)n CH3: 2 (m = 1), 4 (m = 2), 6 (m = 3), 8 (m = 14). The monolayers have been characterized by their pi/A isotherms, their temperature dependence and Brewster angle microscopy (BAM). Amphiphile 1 with the amide group close to the ester group (m = 1) behaves like an unsubstituted fatty acid ester, while 3, 5, and 7, with the amide group in an intermediate and terminal position, exhibit a two-phase region. The amphiphiles 2, 4, 6, and 8, with a reversed orientation of the amide group, all exhibit a two-phase region with higher plateau pressures and lower collapse pressures than those of 1, 3, 5, and 7. For 7 and 8, domains of the liquid condensed (LC) phase are visualized by BAM in the two-phase region. The liquid expanded (LE)/LC-phase transitions are all exothermic with enthalpies deltaH ranging from -31 to -12 kJ/mol. Comparison with other bipolar amphiphiles indicates that the LC phase is better stabilized by the hydroxy and dihydroxy groups than by the amide group. For model compounds of 1-4, optimized conformers in the LE and LC phases have been determined by density functional theory (DFT) calculations.  相似文献   

16.
Whereas there is increasing evidence for ion‐induced protein destabilization through direct ion–protein interactions, the strength of the binding of anions to proteins relative to cation–protein binding has remained elusive. In this work, the rotational mobility of a model amide in aqueous solution was used as a reporter for the interactions of different anions with the amide group. Protein‐stabilizing salts such as KCl and KNO3 do not affect the rotational mobility of the amide. Conversely, protein denaturants such as KSCN and KI markedly reduce the orientational freedom of the amide group. Thus these results provide evidence for a direct denaturation mechanism through ion–protein interactions. Comparing the present findings with results for cations shows that in contrast to common belief, anion–amide binding is weaker than cation–amide binding.  相似文献   

17.
Thin films of fumaramide [2]rotaxane, a mechanically interlocked molecule composed of a macrocycle and a thread in a "bead and thread" configuration, were prepared by vapor deposition on both Ag(111) and Au(111) substrates. X-ray photoelectron spectroscopy (XPS) and high-resolution electron-energy-loss spectroscopy were used to characterize monolayer and bulklike multilayer films. XPS determination of the relative amounts of carbon, nitrogen, and oxygen indicates that the molecule adsorbs intact. On both metal surfaces, molecules in the first adsorbed layer show an additional component in the C 1s XPS line attributed to chemisorption via amide groups. Molecular-dynamics simulation indicates that the molecule orients two of its eight phenyl rings, one from the macrocycle and one from the thread, in a parallel bonding geometry with respect to the metal surfaces, leaving three amide groups very close to the substrate. In the case of fumaramide [2]rotaxane adsorption on Au(111), the presence of certain out-of-plane phenyl ring and Au-O vibrational modes points to such bonding and a preferential molecular orientation. The theoretical and experimental results imply that the three-dimensional intermolecular configuration permits chemisorption at low coverage to be driven by interactions between the three amide functions of fumaramide [2]rotaxane and the Ag(111) or Au(111) surface.  相似文献   

18.
Eliamid is a secondary metabolite isolated from two bacterial strains. This molecule features a linear polyketide backbone terminated by a tetramic acid amide moiety. Among other biological activities, eliamid shows a high and specific cytostatic action on human lymphoma and cervix carcinoma cell lines. The 2,4-anti relative configuration of the C-2,C-4-dimethyl substituted amide fragment was assigned by means of Breit's rule. The absolute configuration of all stereocenters was determined by a combination of degradation methods, structural similarity analysis and total synthesis. The stereogenic centers were introduced by vinylogous Mukaiyama aldol reaction and two consecutive Myers alkylations. The use of pentafluorophenyl ester as acylation agent allowed the efficient formation of tetramic acid amide. The longest linear sequence in the synthesis consist of 13?steps and proceeds with 12?% overall yield. Differential spectroscopy experiments with beef heart submitochondrial particles established that eliamid is a potent inhibitor of the NADH-ubiquinone oxidoreductase complex. Additionally, biosynthesis of eliamid was investigated by feeding experiments with (13) C-labeled precursors.  相似文献   

19.
Aromatic amides bearing 2-azulenyl group on the amide nitrogen were synthesized and their structures were investigated. The π-electron density of the N-aryl group was found to influence the cis-trans conformational preferences of these compounds in solution. X-ray crystallography revealed that the plane of the 2-azulenyl ring has a strong tendency to lie coplanar with the amide plane when the azulene group is located on the same side as the amide oxygen atom.  相似文献   

20.
The metal promoted hydrolysis of nitrile groups in the side chains of tetraazamacrocyclic Cu2+ complexes has been studied by stopped-flow techniques. It is shown that the reaction proceeds by an intramolecular attack of an axially coordinated OH- onto the nitrile group to give the corresponding amide. In alkaline solution the amide then deprotonates and binds to the axial position of the Cu2+ thus preventing further coordination of an OH-. This explains mechanistically that in the Cu2+ complexes of macrocycles carrying two nitrile functions only one is selectively hydrolysed. The nitrile hydrolysis has also been used on a preparative scale to synthesize tetraazamacrocycles with two different side chains. X-Ray diffractions of several products are presented to confirm the structures and the results from the kinetics and equilibria measurements.  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