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.
The synthesis of mimetic of the steroid saponins 1 and 2 was investigated. As a substitute for the complex 22-homo-23-nor-steroid moieties A and B in 1 and 2 diosgenin was introduced. The silyl protected thioorthoester 20 was successfully employed for glucosylation. After selective 2--deacetylation, the glucosylated diosgenyl acceptor 23 was rhamnosylated. The 4---methoxybenzoylated donor 12 gave only minor yields. By using the tri--benzoyl protected donor 15 the [small alpha]--rhamnopyranosyl-(1[rightward arrow]2)-[small beta]--glucopyranosyl-(1[rightward arrow]3[small beta])-diosgenin derivative 25 was obtained.  相似文献   

2.
运用核磁共振技术, 研究了室温离子液体1-丁基-3-甲基咪唑四氟硼酸盐([C4mim][BF4])在重水和氘代氯仿中的聚集行为. 实验结果表明, 随着混合体系中离子液体摩尔分数的增加, 在重水中, 离子液体阳离子上各氢原子的化学位移向低场移动, 且呈现了先急剧变化, 后趋于平缓的变化趋势; 在氘代氯仿中, 离子液体阳离子上H2的化学位移向高场移动, H4和H5以及与氮原子直接相连的甲基和亚甲基上的氢原子的化学位移都向低场移动, 且各氢原子的核磁共振信号发生了变化. 根据质量作用定律及1H NMR化学位移随浓度的变化关系计算了[C4mim][BF4]在重水中的临界聚集浓度和聚集数, 并在离子液体阴、阳离子缔合以及离子液体与溶剂相互作用的基础上对实验结果进行了讨论.  相似文献   

3.
Mono- and diprotonated carbocations and the two-electron oxidation dications derived from parent pyrene and its nonalternant isomers "azupyrene"(dicyclopenta[ef,kl]heptalene)(DCPH) and dicyclohepta[ed,gh]pentalene (DCHP) were studied by DFT at the B3LYP/6-31G(d) level. The most likely site(s) for mono- and diprotonation were determined based on relative arenium ion energies and the structures of the energetically most favored carbocations were determined by geometry optimization. The NMR chemical shifts for the protonated mono- and dications and the oxidation dications were computed by GIAO-NMR at the B3LYP/6-31G(d)//B3LYP/6-31G(d) level and their charge delocalization paths were deduced based on magnitude of the computed [capital Delta][small delta](13)C values and the NPA-derived changes in charges. Relative aromaticity/antiaromaticity in various rings in the energetically favored mono- and dications was estimated via NICS and [capital Delta]NICS. Calculated NMR chemical shift data for and were compared with the available experimental NMR values. The available data on chemical and physical properties of DCPH and DCHP are extremely limited and biological activity data are non-existent. The present study provides the first glance into their carbocations and oxidation dications, while augmenting and reinforcing the previous stable ion data on the pyrenium cations.  相似文献   

4.
Proton NMR spectra of short peptides with a glycyl (Gly) or N-methylglycyl (sarcosyl, Sar) residue were measured in various mixed solvents with a wide range of dielectric constants: 78.3-2.3. From analyses of the octet and quartet signals of the geminal alpha-protons of Gly and Sar residues, respectively, we have estimated chemical shift differences between the two alpha-protons, Delta delta(alpha/alpha'). It is found that the Delta delta(alpha/alpha') values increase with decreasing solvent polarity and the increasing rates depend significantly on amino acid sequences. By referring to infrared spectra and chemical shift of the terminal NH protons, deltaNH, of the peptides, the Delta delta(alpha/alpha') values were found to be a good probe of beta-turn formation. From solvent-dependent change of Delta delta(alpha/alpha'), we estimated the free energies for the beta-turn formation and compared the results with those estimated from deltaNH. Using the resulting free energies, we have discussed effects of solvent on the beta-turn formation.  相似文献   

5.
It has been suggested that astaxanthin (3,3'-dihydroxy-beta,beta-carotene-4,4'-dione) in the carotenoprotein alpha-crustacyanin occurs in the diprotonated form. As a model system for protonated astaxanthin in [small alpha]-crustacyanin the reactions of canthaxanthin ([small beta],[small beta]-carotene-4,4[prime or minute]-dione) with Bronsted acids (CF(3)COOH and CF(3)SO(3)H) and the Lewis acid BF(3)-etherate have been investigated. Structures of C-5 protonated, C-7 protonated, enolised O-4 protonated and O-4,4[prime or minute], C-7 triprotonated canthaxanthin have been established by VIS-NIR and NMR spectroscopy. The charge distribution in the cations has been considered by comparison of the (13)C chemical shift difference relative to neutral relevant carotenoid models. The experimental evidence for protonated canthaxanthins differs significantly from previous AM1 calculations. Experimental data for O-4,4[prime or minute], C-7 triprotonated canthaxanthin relative to C-7 protonated canthaxanthin is considered a relevant model for O-4,4[prime or minute] diprotonated canthaxanthin, in comparison with neutral canthaxanthin. The positive charge was mainly located at C-6/6[prime or minute][dbl greater-than] C-8/8[prime or minute] > C-10/10[prime or minute] > C-12/12[prime or minute] > C-14/14[prime or minute][similar] C-15/15[prime or minute] in the polyene chain. Moreover, it was inferred that only 14% of the positive charge is delocalised to the polyene chain, the remaining charge must therefore be located at the protonated carbonyl moiety. The results are discussed in relation to previous solid state NMR studies of (13)C labelled astaxanthin in [small alpha]-crustacyanin and recent X-ray analysis of [small beta]-crustacyanin.  相似文献   

6.
We report on a spectrophotometric kinetic study of the effect of Li(+) and K(+) cations on the ethanolysis of 4-nitrophenyl dimethylphosphinate () in ethanol at 25 [degree]C. The nucleophilic displacement reaction of with LiOEt and KOEt in the absence and presence of 18-crown-6 ether (18-C-6) furnished observed first-order rate constants which increase in the order EtO(-) < KOEt < LiOEt. The kinetic data are analyzed in terms of a scheme which assigns concurrent kinetic activity to free ethoxide and metal alkoxide, to obtain the second-order rate coefficients for reaction of the metal ion-ethoxide pairs, k(MOEt). Derived [small delta]G(ip), [small delta]G(ts) and [capital Delta]G(cat) values quantify ground state and transition state stabilization by the metal ions to give [small delta]G(ts) > [small delta]G(ip) for Li(+) and [small delta]G(ts)[similar][small delta]G(ip) for K(+). These results indicate moderate catalysis by Li(+), with manifesting lesser susceptibility to catalysis than other substrates previously studied. Second-order rate constants for the reaction of the aryl dimethylphosphinates with free EtO(-) were obtained from plots of log k(obs)vs. [KOEt], measured in the presence of excess 18-C-6. Hammett plots with [sigma] and [sigma][degree] substituent constants give significantly better correlation of rates than [sigma](-) and yield a moderately large [small rho]([small rho][degree]) value; this is interpreted in terms of a stepwise mechanism involving rate-limiting formation of a pentacoordinate intermediate. Comparison of the present results with those of Williams on the aqueous alkaline hydrolysis of Me(2)P(O)-OPhX and Ph(2)P(O)-OPhX esters, establishes the rationale for a change in mechanism in the more basic EtO(-)/EtOH nucleophile/solvent system by a stepwise mechanism instead of a concerted one in aqueous base. Structure-reactivity correlations following Jencks show that the change in mechanism is accounted for by cross interactions between the nucleophile and the leaving group in the transition state. The observed duality of mechanism is rationalized on the basis of the More O'Ferrall-Jencks diagram, as a spectrum of transition states covering a wide range of nucleophile and leaving group basicities.  相似文献   

7.
The 1H chemical shifts of 124 compounds containing a variety of functional groups have been recorded in CDCl3 and DMSO-d6 (henceforth DMSO) solvents. The 1H solvent shift Delta delta = delta(DMSO) - delta(CDCl3) varies from -0.3 to +4.6 ppm. This solvent shift can be accurately predicted (rms error 0.05 ppm) using the charge model of alpha, beta, gamma and long-range contributions. The labile protons of alcohols, acids, amines and amides give both, the largest solvent shifts and the largest errors. The contributions for the various groups are tabulated and it is shown that for H.C.C.X gamma-effects (X = OH, NH, =O, NH.CO) there is a dihedral angle dependence of the gamma-effect. The group contributions are discussed in terms of the possible solvent-solute interactions. For protic hydrogens, hydrogen bonding is the dominant interaction, but for the remaining protons solvent anisotropy and electric field effects appear to be the major factors.  相似文献   

8.
The dependence of the 129Xe NMR chemical shift value of XeF2 on temperature and concentration was determined in a variety of prototypic media: in acidic (anhydrous HF, aHF), nonprotic but polar (dichloromethane), and basic (CD3CN-EtCN, 1:3 v/v) solvents. The 129Xe NMR spectra of a representative series of organoxenon(II) salts [RXe][Y] (R = C6F5, heptafluoro-1,4-cyclohexadien-1-yl (cyclo-1,4-C6F7), pentafluoro-1,4-cyclohexadien-3-on-1-yl (cyclo-3-O-1,4-C6F5), CF2=C(CF3), (CF3)2CFC[triple bond]C, C4H9C[triple bond]C; Y = BF4, AsF6) in aHF showed, compared with XeF2-aHF, a quantitatively less distinct but qualitatively related dependence of delta(129Xe) vs temperature. The dependence of their delta(129Xe) values on concentration in aHF is negligible. An explanation for the different behavior of [RXe][Y] and XeF2 is offered.  相似文献   

9.
The complexes [MeBu 3N] 2[UCl 6] and [BuMe 2Im] 2[UCl 6] were characterized in the solid state and in solution of [MeBu 3N][Tf 2N], [BuMe 2Im][Tf 2N], and [BuMeIm][Tf 2N] room-temperature ionic liquids using single-crystal XRD, EXAFS, electrochemistry, UV-visible absorption spectroscopy, and NMR. In the solid state and in solution, the existence of hydrogen bonding between the UCl 6 (2-) complex and the ionic liquid cations was revealed by these techniques. The MeBu 3N (+) cation interacts with UCl 6 (2-) via the protons on the alpha-carbon atoms of nitrogen. The protons of the imidazolium ring account for the interaction between the BuMe 2Im (+) cation and the UCl 6 (2-) anion. For the BuMeIm (+) cation the major interaction was confirmed between the most acidic proton on C(2) and the chlorides of UCl 6 (2-). The experimental results also show that the intensity of the interaction between the UCl 6 (2-) anion and the cation varies with the ionic liquid cation in the following order: MeBu 3N (+) approximately BuMe 2Im (+) < BuMeIm (+).  相似文献   

10.
The chemical shifts, temperature coefficients and inter-residual rotating-frame Overhauser effect (ROE)s for the hydroxy protons of some alpha-(1,2)-, alpha-(1,3)- and alpha-(1,6)-linked di- and trimannosides have been measured for samples in 85% H2O/15% acetone-d6 solution. These mannosides, Manalpha(1-->2)ManalphaOMe (1) Manalpha(1-->3)ManalphaOMe (2), Manalpha(1-->6)ManalphaOMe (3), Manalpha(1-->2)Manalpha(1-->2)ManalphaOMe (4), Manalpha(1-->2)Manalpha(1-->3)ManalphaOMe (5), Manalpha(1-->2)Manalpha(1-->6)ManalphaOMe (6) and Manalpha(1-->3)[Manalpha1-->6]ManalphaOMe (7), are substructures of the N-glycan Man-9.The NMR data show that the hydration of each individual hydroxyl group in the di- and trisaccharides is very similar to the hydration of the corresponding hydroxyl in the monomeric methyl alpha-D-mannoside. No hydrogen-bond interactions were found to stabilize the conformations of the alpha-(1,2)- and alpha-(1,6)-linkages and the chemical shifts for the hydroxy proton resonances of the alpha-(1,6)-linkage indicated high-conformational flexibility. For the alpha-(1,3)-linkage, however, the downfield shift for the signal of O(2)H of the 3-substituted residue together with the ROE between this proton and H5' on the next residue suggest some weak inter-residue interactions.  相似文献   

11.
The NMR spectra of the benzohexahelicenes have been analyzed. The conformation of the helical structure is not changed to a large extent by the benzo-groups. Some flattening of a part of the helix results in a small downfield shift of the protons on the terminal rings and in a larger upfield shift of proton A in II ascribed to a change in the ring current in the pitch of the helix. A gradual upfield shift is found for protons of the terminal rings by an increase in overcrowding of the benzo-groups.  相似文献   

12.
The homogeneous system of 12-tungstophosphoric acid and diethylene glycol was studied using IR and NMR spectra. It was found that the protons in 12-tungstophosphoric acid formed proton oxonium ions with the hydroxyl oxygen in diethylene glycol by hydrogen-bonds, and the formed proton oxonium ions could react with the terminal oxygens of heteropoly anions. The dehydration-cyclization mechanism of diethylene glycol in the presence of heteropoly acid was also proposed.  相似文献   

13.
[structure: see text] The conventional interpretation of proton NMR chemical shifts is supported by large basis set ab initio quantum mechanical calculations. The benzene protons are predicted to lie within the deshielding zone defined in terms of the out-of-plane magnetic shielding domain. However, ring currents by themselves are not sufficient to account quantitatively for the observed benzene proton downfield chemical shift. sigma-Electron contributions must also be taken into account. The conventional explanation for the ethyne proton chemical shift is valid.  相似文献   

14.
18-冠-6与Eu(fod)3+在CCl4中的化学平衡及其配合物的结构   总被引:1,自引:0,他引:1  
在室温下测定了CCl_4溶液中,18-冠-6的亚甲基、Eu(fod)的t—丁基质子的PMR化学位移与溶液组成的相互关系。提出了生成配合物RS,R_2S,R_2,R_3(S为18-冠-6,R为Eu(fod)_3)的模式。求得了平衡常数值(升/摩):K_1=[RS]/([R][S])=2.O×10~5,K_2=[R_2S]/([R][RS])=1.0×10~3,Q_1=[R_2]/[R]~2=4.2×10~2,Q_2=[R_3]/([R_2][R])=4.0×10~1。并测得在配合物RS,R_2S中18-冠-6质子的平均化学位移分别为4.90,10.84(ppm,相对于纯18-冠-6); 在配合物R,R_2,R_3,RS,R_2中t-丁基质子平均化学位移分别为1.53,2.33,0.54,1.02,2.42(ppm,相对于TMS)。 根据配合物的化学位移值,并考虑分子及分子轨道的对称性、位阻等因素提出了配合物R_2,RS,R_2S的可能的结构模型。  相似文献   

15.
Aromatic stacking of 9,9'-(alpha,omega-alkanediyl)bis[adenine] (1), 1,1'-(alpha,omega-alkanediyl)bis[benzimidazole] (2), and 9-[omega-(benzimidazol-1-yl)alkyl]adenine (3) were studied at low concentrations of these compounds by means of UV and NMR spectroscopies. The UV hypochromic effect at T degrees C was determined as the ratio of the integration strength B at T degrees C (T = 27, 40, and 50) to that at 60 degrees C. The UV hypochromic effects of 1 and 3 were remarkable in water, suggesting a formation of intramolecular aromatic stacking, while the UV data of 2 did not present unambiguous evidence supporting aromatic stacking. A difference of chemical shift of each aromatic ring proton between 27 and 80 degrees C, that is Deltadelta = delta(80 degrees C) - delta(27 degrees C), was given as an indication of the aromatic stacking in the NMR study. On the basis the data of Deltadelta, 1 and 3 were stabilized by a stacking interaction in the buffer solution at pD 7.0 but not in the organic solvents. On the other hand, the NMR data did not indicate the formation of aromatic stacking of 2 either in the organic solvents or in the aqueous solution. The thermodynamic parameters of the intramolecular aromatic stacking of 3 were determined by means of NMR spectroscopy.  相似文献   

16.
[structures: see text] The absolute configuration of 1,2-, 1,3-, 1,4-, and 1,5-diols formed by two secondary (chiral) hydroxy groups can be deduced by comparison of the NMR spectra of the corresponding bis-(R)- and bis-(S)-MPA esters. The correlation between the NMR spectra of the bis-ester derivatives and the absolute stereochemistry of the diol involves the comparison of the chemical shifts of the signals for substituents R1/R2 and for the hydrogens attached to the two chiral centers [H(alpha)(R1) and H(alpha)(R2)] in the bis-(R)- and the bis-(S)-ester and is expressed as delta deltaRS. Theoretical calculations [energy minimization by semiempirical (AM1), ab initio (HF), DFT (B3LYP), and Onsager methods, and aromatic shielding effect calculations] and experimental data (NMR and CD spectroscopy) indicate that in these bis-MPA esters, the experimental delta deltaRS values are the result of the contribution of the shielding/deshielding effects produced by the two MPA units that combine according to the actual stereochemistry of the diol. The reliability of these correlations is demonstrated with a wide range of diols of known absolute configuration derivatized with MPA and 9-AMA as auxiliary reagents. A simple graphical model that allows the simultaneous assignment of the two asymmetric carbons of a 1,n-diol by comparison of the NMR spectra (delta deltaRS signs) of its bis-(R)- and bis-(S)-AMAA ester derivatives is presented.  相似文献   

17.
The synthesis of [2.2]paracyclophane/dehydro[14]annulene hybrids 1 and 2 is reported. Comparison of the proton NMR spectra of 1 and 2 with their open precursors and with related model compounds reveals the pronounced effect of macrocycle formation upon the cyclophane protons H15/H16, which lie above the shielding cone of the diatropic [14]annulene moiety. [structure: see text]  相似文献   

18.
The diagnostic values of the following three spectral criteria for the configuration of N-acyl-alpha,beta-dehydro-alpha-amino acid esters were examined: (i) the proton at the beta-position at the double bond of a Z-isomer is shielded if compared with the respective E-isomer (delta(beta)Z < delta(beta)E); (ii) the proton at the nitrogen atom is shielded in a Z-isomer in comparison with the corresponding E-isomer (delta(NH)Z < delta(NH)E); and (iii) changing of the solvent from CDCl3 to deuterated trifluoroacetic acid (TFA) causes shielding of the H(beta) vinylic proton of an E-isomer or deshielding of the respective proton of the Z-isomer (delta(CDCl3)E > delta(TFA)E or delta(CDCl3)Z < delta(TFA)Z). The investigations were based on a set of 22 (Z)- and (E)-N-acyl-alpha,beta-dehydro-alpha-amino acid esters of diverse structures, with aliphatic, aromatic and heteroaromatic substituents at the vinylic beta-carbon; most of the examined compounds were hitherto unknown. The application of the substituent effect additivity rule given by Pascual et al. for olefinic protons leads to evidently erroneous configuration assignments of N-acyl-alpha,beta-dehydro-alpha-amino acid esters. The considered criteria were fulfilled for all the examined cases with one exception [the second criterion for the alpha-pivaloylamino-beta-(2-furyl)acrylates]. The comparison of changes in the chemical shifts of H(beta) vinylic protons in CDCl3 and deuterated TFA seems to be the most reliable and useful configuration criterion, as it can be used in the case of a single isomer.  相似文献   

19.
A new functional group, the hydroxy group, was inserted into a Betti base by reaction with salicylaldehyde, and the naphthoxazine derivatives thus obtained were converted by ring-closure reactions with formaldehyde, acetaldehyde, propionaldehyde or phosgene to the corresponding naphth[1′2′:5,6][1,3]oxazino[3,2-c][1,3]benzoxazine derivatives. Further, the conformational analysis of these polycyclic compounds by NMR spectroscopy and an accompanying molecular modelling are reported; especially, both quantitative anisotropic ring current effects of the aromatic moieties in these compounds and steric substituent effects were employed to determine the stereochemistry of the naphthoxazinobenzoxazine derivatives.  相似文献   

20.
Biginelli reaction of thiourea, 2‐hydroxy‐1‐naphthaldehyde, and acetoacetic ester (or benzoyl acetone) under solvent‐free conditions and MW irradiation gave novel 3‐thioxo‐2,3,4,5‐tetrahydro‐1H‐1,5‐methanonaphtho[1,2‐g][1,3,5]oxadiazocine derivatives. Subsequent reaction of the obtained compounds with α‐chloroacetamide led to 5‐methyl‐5H,13H‐5,13‐methanonaphtho[1,2‐g] thiazolo[2,3‐d][1,3,5]oxadiazocin‐1(2H)‐ones, which were converted to the Z‐isomers of 2‐arylylidene‐5H,13H‐5,13‐methanonaphtho[1,2‐g]thiazolo[2,3‐d][1,3,5]oxadiazocin‐1(2H)‐one derivatives by reaction with arylaldehydes. The structures of the products were characterized by 1H NMR, 13C NMR spectra, and X‐ray analysis.  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