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 109 毫秒
1.
The 1H-NMR spectra of 2-(nitromethylidene)pyrrolidine ( 7 ), 1-methyl-2-(nitromethylidene)imidazolidind ( 10 ) and 3-(nitromethylidene)tetrahydrothiazine ( 11 ) in CDCl3 and (CD3)2SO indicate that these compounds have the intramolecularly H-bonded structures (Z)- 7 , (E)- 10 and (Z)- 11 while the N-methyl derivative 8 of 7 is (E)-configurated in both solvents. 1-Benzylamino-1-(methyltio)-2-nitroehtylene ( 13 ), an acylic model, has the H-bonded configuration (E)- 13 in CDCl3 and in (CD3)2SO. 2-(Nitromethylidene)thiazolidine ( 3 ) has the (E)-configuration in CDCl3 but exists in (CD3)2SO as a mixture of (Z)- and (E)-isomers with the former predominating. Both species are detected to varying proportions in a mixture of the two solvents. 15N-NMR spectroscopy of 3 ruled out unambiguously the nitronic acid structure 6 and the nitromethyleimine structure 5 . The N-methyl derivative 4 of 3 is (Z)-configurated in (CD3)2SO. Comparison of the olefinic proton shifts of (Z)- 3 and (Z)- 4 with those of analogues and also of 1,1-bis(methylti)-2-nitroethylene ( 12 ) shows decreased conjugation of the lone pair of electrons of the ring N-atom in (Z)- 3 and (Z)- 4 . This is also supported by 13C-NMR studies. Plausible explanations for the phenomenon are offered by postulating that the ring N-atoms are pyramidal in (Z)- 3 and (Z)- 4 and planar in other cases or, alternatively, that the conjugated nitroenamine system gets twisted due to steric interaction between the NO2-group and the ring S-atom. Single-crystal X-ray studies of 3 and 8 show that the former exists in the (Z)-configuration and the latter in (E)-configuration; the ring N-atom in the former has slightly more pyramidal character than in the latter.  相似文献   

2.
1D 1H and 1 3C, and 2D 1H NOESY NMR were used to establish that cyanoacetylhydrazones of aliphatic and aromatic aldehydes and ketones in CDCl3 and (CD3)2SO solutions have a linear structure, exist as a mixture of geometric isomers by the C = N bond, and exhibit hindered amide rotation.  相似文献   

3.
INDOR experiments indicate 4J(P? C? C? O? H) to be positive for the β-hydroxyphosphonate ester RS,RS dissolved in (CD3)2CO or (CD3)2SO. The higher value observed in (CD3)2SO than in (CD3)2CO shows that a W geometry makes this coupling constant more positive.  相似文献   

4.
Summary The1H and13C NMR spectra of the lupin alkaloidangustifoline 1 in four solvents (cyclohexane-d12, CDCl3, CD3CN, and C6D6) were assigned using 2D H,H and H,C COSY and 2D J-resolved spectra. The torsional HCCH angles calculated from the vicinalJ HH coupling constants are essentially in agreement with those expected for the deformed all-chair conformation withendo oriented N(12)-H bond, reported earlier for1 in the solid state. Some arguments seem to point, however, to a small contribution of other conformations: with ring A deformed in another direction, deformed all-chair withexo oriented N(12)-H bond and/or a conformation with ring C in the boat form.Lupin Alkaloids, part 7  相似文献   

5.
The 1H and 13C NMR spectra of 10‐deoxymethynolide (1), 8.9‐dihydro‐10‐deoxymethynolide (2) and its glycosylated derivatives (3–9) were analyzed using gradient‐selected NMR techniques, including 1D TOCSY, gCOSY, 1D NOESY (DPFGSENOE), NOESY, gHMBC, gHSQC and gHSQC‐TOCSY. The NMR spectral parameters (chemical shifts and coupling constants) of 1–9 were determined by iterative analysis. For the first time, complete and unambiguous assignment of the 1H NMR spectrum of 10‐deoxymethynolide (1) has been achieved in CDCl3, CD3OD and C6D6 solvents. The 1H NMR spectrum of 8,9‐dihydro‐10‐deoxymethynolide (2) was recorded in CDCl3, (CD3)2CO and CD3OD solutions to determine the conformation. NMR‐based conformational analysis of 1 and 2 in conjugation with molecular modeling concluded that the 12‐membered ring of the macrolactones may predominantly exist in a single stable conformation in all solvents examined. In all cases, a change in solvent caused only small changes in chemical shifts and coupling constants, suggesting that all glycosylated methymycin analogs exist with similar conformations of the aglycone ring in solution. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
Epoxy derivatives of internal fluoroolefins (both cis and trans isomers) react in a stereospecific manner with (1S)- or rac-camphor thiosemicarbazone in a polar aprotic medium to give the corresponding 5-fluoro-4-hydroxy-4,5-bis(polyfluoroalkyl)thiazol-2-ylhydrazones. From unsymmetrical 2,3-epoxydodecafluorohexane a mixture of regioisomeric hydrazones is formed. According to the 1H and 1 9F NMR data, the resulting trans-hydrazones in (CD3)2SO and CDCl3 exist as mixtures of diastereoisomers occurring in a dynamic equilibrium.  相似文献   

7.
Reaction of Zn(OAc)2 ? 2 H2O with 2,6‐diisopropylphenyl phosphate (dippH2) in the presence of pyridine‐4‐carboxaldehyde (Py‐4‐CHO) in methanol resulted in the isolation of a tetrameric zinc phosphate cluster [Zn(dipp)(Py‐4‐CH(OH)(OMe))]4 ? 4 MeOH ( 1 ) with four hemiacetal moieties stabilized on the double‐4‐ring inorganic cubane cluster. The change of solvent from methanol to acetonitrile leads to the formation of [Zn(dipp)(Py‐4‐CHO)]4 ( 2 ), in which the coordinated Py‐4‐CHO retains its aldehydic form. Dissolution of 1 in CD3CN readily converts it to the aldehydic form and yields 2 . Similarly 2 , which exists in the aldehyde form in CD3CN, readily converts to the hemiacetal form in CD3OD/CH3OH. Compound 1 is an unprecedented example in which four hemiacetals have been stabilized on a single molecule in the solid state retaining its stability in solution as revealed by its 1H NMR spectrum in CD3OD. The solution stability of 1 and 2 has further been confirmed by ESI‐MS studies. To generalize the stabilization of multiple hemiacetals on a single double‐four‐ring platform, pyridine‐2‐carboxaldehyde (Py‐2‐CHO) was used as the auxiliary ligand in the reaction between zinc acetate and dippH2, leading to isolation of [Zn(dipp)(Py‐2‐CH(OH)(OMe))]4 ( 3 ). Understandably, recrystallization of 3 from acetonitrile yields the parent aldehydic form, [Zn(dipp)(Py‐2‐CHO)]4 ( 4 ). Single‐crystal X‐ray diffraction studies reveal that supramolecular bonding, aided by hydrogen‐bonding interactions involving the hemiacetal functionalities (C?OH, C?OMe, and C?H), are responsible for the observed stabilization. The hemiacetal/aldehyde groups in 1 and 2 readily react with p‐toluidine, 2,6‐dimethylaniline, and 4‐bromoaniline to yield the corresponding tetra‐Schiff base ligands, [Zn(dipp)(L)]4 (L=4‐methyl‐N‐(pyridin‐4‐ylmethylidene)aniline ( 5 ), 2,6‐dimethyl‐N‐(pyridin‐4‐ylmethylene)‐aniline ( 6 ), and 4‐bromo‐N‐(pyridin‐4‐ylmethylene)aniline ( 7 )). Isolation of 5 – 7 opens up further possibilities of using 1 and 2 as new supramolecular synthons and ligands.  相似文献   

8.
F. Baert  R. Fouret  M. Sliwa  H. Sliwa 《Tetrahedron》1980,36(19):2765-2774
IR, 1H NMR and X-ray crystallographic studies of the title compounds show that the trans-isomer exhibit a diaxially substituted half-chair conformation of the dihydropyran ring owing to destabilization of the alternative overcrowded diequatorial conformation Steric hindrance prevents a closer approach of the hydroxyl group to the ring oxygen limiting the strength of the OH…O intramolecular hydrogen bonding and allowing thermal vibration of large amplitude. The cis isomer adopts a C(3) sofa conformation for the dihydropyran ring, tightly bonded by a strong OH…π intramolecular hydrogen bonding which contributes to a higher rigidity of the whole molecule.  相似文献   

9.
Conformational changes of amide cavitands A – C were investigated at varied temperatures and in several solvents. While cavitands A and B , with comparatively smaller substituents such as Et and iPr, were always in vase conformation in non‐polar solvents such as CDCl3, CD2Cl2, (D8)THF, and C6D6, their thermoswitching (vase to kite) was observed in polar solvents such as (D7)DMF and (D6)DMSO or in the presence of acid (TFA) and H‐bonding inhibitor (TFE). Intra‐ and interannular H‐bonds of A and B were clearly observed by low‐temperature 1H‐NMR spectra in CDCl3. No conformational change of cavitand C with bigger substituent (tBu) was observed under any tested temperature range and in polar or non‐polar solvents; C was always in the kite conformation.  相似文献   

10.
The Cinchona alkaloid analogs (+)- and (?)- 5 with a quinuclidine-2-methanol residue attached to C(2) of a 9,9′-spirobifluorene moiety were prepared as a racemic mixture by reacting lithiated 2-bromo-9,9′-spirobifluorene 7 with (2-ethoxycarbonyl)quinuclidine (±)- 6 to give ketone (±)- 8 , followed by diastereoselective reduction with diisobutylaluminum hydride (DIBAL-H). The absolute configuration at C(9) and C(8), i.e., at the methanol bridge and the adjacent quinuclidine C-atom, in the two enantiomers of 5 is identical to the configuration at the corresponding centers in (?)-quinine ( 1 ) and (+)-quinidine ( 2 ), respectively. For the optical resolution of (±)- 5 , a chiral stationary phase for HPLC was prepared by covalently bonding quinine via a thiol spacer to a silica-gel surface. The enantiomer separation was accomplished at an α value of 1.61 with (±)- 5 being eluted last, in agreement with 1H-NMR studies in CDCl3 which showed that (+)- 5 underwent a more stable host-guest association with quinine than (?)- 5 . 1H{1H} Nuclear Overhauser effect (NOE) difference spectroscopical analysis of the host-guest associations with quinine in CDCl3, combined with computer-model examinations, allowed the assignment of the absolute configurations as (+)-(8R,9S)- 5 and (?)-(8S,9R)- 5 . A detailed conformational analysis displayed excellent agreement between the results of computational methods (Monte Carlo multiple minimum simulations, analyses of the total energy as a function of the flexible dihedral angles in the molecule) and 1H{1H}-NOE difference spectroscopical data. It was found that (?)- 5 and (+)- 5 differ significantly in their conformational preference from their natural counterparts quinine ( 1 ) and quinidine ( 2 ). Whereas the natural alkaloids prefer the ‘open’ conformation, with the quinuclidine N-atom pointing away from the quinoline ring, analog (±)- 5 adopts preferentially (by ca. 4 kcal mol?1) a ‘closed’ conformation, in which the quinuclidine N-atom points into the cleft of the 9,9′-spirobifluorene moiety. Since the basic quinuclidine N-atom in the ‘closed’ conformation is sterically shielded from forming strong H-bonds, the new Cinchona alkaloid analogs form less stable host-guest associations via H-bonding than quinine or quinidine.  相似文献   

11.
According to the 1H, 13C, and 19F NMR data, fluoroalkyl-containing 1,2,3-trione 2-arylhydrazones in CDCl3 exist exclusively, while in (CD3)2CO preferentially, as isomers in which the acyl or aroyl group is involved in intramolecular hydrogen bond. The isomer structure was assigned on the basis of the chemical shifts of the carbonyl carbon atoms and fluorine atoms and carbon-fluorine spin-spin coupling constants J C-F. X-Ray diffraction data showed that 1,2,3-trione 2-arylhydrazones in crystal have the same structure as in CDCl3 solution. Quantum-chemical calculations were performed to rationalize predominant formation of 1,2,3-trione 2-arylhydrazone isomers with a free polyfluoroacyl group.  相似文献   

12.
The electron polarizabilities (α0·1024/cm3 molec.−1) were estimated from the data on refractive indices and molar volumes of H/D isotopomers of methanol at 25 °C using the Lorentz-Lorentz formula: 3.265 (CH3OH), 3.260 (CH3OD), 3.235 (CD3OH), and 3.231 (CD3OD). A relationship between the isotope effects for α0 and volume (packing) changes in the structure of liquid methanol induced by deuterosubstitution in the methanol molecule was proposed. __________ Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 8, pp. 1927–1928, August, 2005.  相似文献   

13.
The ion pair of the stereolabile C3‐symmetric, i+o proton complex [ 1? H]+ of diaza‐macropentacycle 1 and the configurationally stable Δ‐TRISPHAT ([Δ‐ 3 ]?) anion exists in the form of two diastereomers, namely, [Δ‐( 1? H)][Δ‐ 3 ] and [Λ‐( 1? H)][Δ‐ 3 ], the ratio of which, in terms of diastereomeric excess (de) decreases in the order [D8]THF (28 %)>CD2Cl2 (22 %)>CDCl3 (20 %)>[D8]toluene (16 %)>C6D6 (7 %)>[D6]acetone (0 %) at thermodynamic equilibrium. Except in the case of [D6]acetone, the latter is reached after a period of time that increases from 1 h ([D8]THF) to 24 h (CDCl3). Moreover, the initial value of the de of [ 1? H][Δ‐ 3 ] in CDCl3, before the thermodynamic equilibrium is reached, depends on the solvent in which the sample has been previously equilibrated (sample “history”). This property has been used to show that the crystals of [ 1? H][Δ‐ 3 ] formed by slow evaporation of CH2Cl2/CH3OH mixtures had 100 % de, which indicates that [ 1? H][Δ‐ 3 ] has enjoyed a crystallization‐induced asymmetric transformation. Structural studies in solution (NMR spectroscopy) and in the gas phase by calculations at the semiempirical PM6 level of theory suggest that the optically active anion is docked on the i+ (endo) external side of the proton complex such that one of the aromatic rings of [Δ‐ 3 ]? is inserted into a groove of [ 1? H]+, a second aromatic ring being placed astride the outside i+ pocket. Solvent polarity controls the thermodynamics of inversion of the [ 1? H]+ propeller. However, both polarity and basicity control its kinetics. Therefore, the rate‐limiting steps correspond to the ion‐pair separation/recombination and [ 1? H]+/ 1 deprotonation/protonation processes, rather than the inversion of [ 1? H]+, the latter being likely to take place in the deprotonated form ( 1 ).  相似文献   

14.
The determination of the enantiomeric composition of chiral compounds by1H,13C, and31P NMR spectroscopy in the presence of (S)-1,1′-binaphthyl-2,2′-diol demonstrates that the enantioselectivity of the method increases when the polarity of a solvent decreases as follows: CD3OD-D2O (4 ∶ 1) < CD3OD < CDCl3 < CDCl3-CCl4 (1 ∶ 1) < C6D6. The effect is caused by increase in stability of solvating agent-substrate complexes formed through the hydrogen bonds. Pantolactone, esters of substituted cyclopropanecarboxylic acids, amino alcohol propranolol, and 2,2′-bis(diphenylphosphinyl)-1,1′-binaphthyl were used as the substrates.  相似文献   

15.
The keto-enol equilibrium of 2 × 10−3 M solutions of (acetoacetyl)ferrocene and 1,1′-bis(acetoacetyl)ferrocene was studied by 1H NMR spectroscopy in a series of aprotic solvents such as DMSO-d 6, (CD3)2CO, CDCl3, CD2Cl2, CCl4, and C6D6 at a temperature of 20°C. It was established that the calculated enolization constants increase with a decrease in the polarity of solvent molecules. The results complement and combine known empirical rules about the effect of the medium on keto-enol equilibria.  相似文献   

16.
The discharge-flow method with resonance fluorescence detection of OH radicals was applied to obtain the rate constant value of k D = 1.95 ± 0.14 (1σ) 1010 cm3 mol-1s-1 at 298 K. Combination with k H from our previous study gives the kinetic isotope effect of k H / k D = 5.33 ± 0.41. OH + CH3C(O)CH3 → Products (H) OH + CD3C(O)CD3 → Products(D) This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

17.
High resolution proton decoupled 13C and 31P n.m.r. spectra of bis(diphenylphosphino)acetylene in (CD3)2SO and CDCl3 are analysed as ABX spectra to give the relative chemical shifts of the 13C and 31P nuclei as well as the spin–spin coupling constants 3J(PP) and nJ(PC). The differences in 31P shieldings are due to secondary 13C isotope effects which have been observed to be negligible over more than two bonds.  相似文献   

18.
NMR spectra of the synthesized azo dyes, 5‐arylazo‐pyrimidine (1H,3H,5H)‐2,4,6‐triones (5a–g), 1,3‐dimethyl‐5‐arylazo‐pyrimidine (1H,3H,5H)‐2,4,6‐triones (6a–g), and 5‐arylazo‐2‐thioxo‐pyrimidine (1H,3H,5H)‐4,6‐diones (7a–g) were studied in (CD3)2SO (three drops of CD3OD were added into solutions of the dyes in two different concentrations). All dyes showed intramolecular hydrogen bonding. Dyes 5a–7a showed bifurcated intramolecular hydrogen bonds. Tautomeric behaviours of some of N‐methylated azo dyes (6a‐g) were studied in two different concentrations. The solvent–substrate proton exchange of dyes 5a–d, 6a and 7a–e was examined in presence of three drops of CD3OD. The dyes which were soluble in (CD3)2SO containing CD3OD showed isotopic splitting (β‐isotope effect) in the 13C NMR spectra. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
The 1H and 13C{1H} chemical shifts and 1H spin–spin couplings of sulfur mustards, nitrogen mustards, and lewisites scheduled in the Chemical Weapons Convention, and those of bis(2‐chloromethyl)disulfide, were determined in CDCl3, CD2Cl2, and (CD3)2CO. Accurate parameters of this kind of series can be used for evaluating the current molecular modeling programs and the chemical shift and coupling constant prediction possibilities of the programs. Several prediction tests were made with commercial programs, and the results are reported here. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

20.
The possible stable forms and molecular structures of 1‐cyclohexylpiperazine (1‐chpp) and 1‐(4‐pyridyl)piperazine (1‐4pypp) molecules have been studied experimentally and theoretically using nuclear magnetic resonance(NMR) spectroscopy. 13C, 15N cross‐polarization magic‐angle spinning NMR and liquid phase1H, 13C, DEPT, COSY, HETCOR and INADEQUATE NMR spectra of 1‐chpp (C10H20N2) and 1‐4pypp (C9H13N2) have been reported. Solvent effects on nuclear magnetic shielding tensors have been investigated using CDCl3, CD3 OD, dimethylsulfoxide (DMSO)‐d6, (CD3)2CO, D2O and CD2Cl2. 1H and 13C NMR chemical shifts have been calculated for the most stable two conformers, equatorial–equatorial (e–e) and axial–equatorial (a–e) forms of 1‐chpp and 1‐4pypp using B3LYP/6‐311++G(d,p)//6‐31G(d) level of theory. Results from experimental and theoretical data showed that the molecular geometry and the mole fractions of stable conformers of both molecules are solvent dependent. Copyright © 2009 John Wiley & Sons, Ltd.  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