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 31 毫秒
1.
An effective pulse sequence for measuring H–H coupling constants, named BASHD‐J‐resolved‐COSY, has been developed. In the spin systems such as –CHA–CHB(CH3)–CHC–, a methine proton HB splits into a multiplet owing to several vicinal couplings, resulting in attenuation of its cross‐peak intensity. Therefore, the measurements of 3JH–H with respect to HB are generally difficult in the E‐COSY‐type experiments. With the aim of accurate measurements of 3JH‐H in such a spin system, we have developed a new pulse sequence, which selectively decouples the secondary methyl group. The proposed pulse sequence provides the simplified cross‐peak patterns, which are suitable for reliable measurements of 3JH‐H in a complicated natural product. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

2.
Natural products often possess various spin systems consisting of a methine group directly bonded to a methyl group (e.g. –CHa–CHb(CH3)–CHc–). The methine proton Hb splits into a broadened multiplet by coupling with several vicinal protons, rendering analysis difficult of nJC–H with respect to Hb in the J‐resolved HMBC‐1. In purpose of the reliable and easy measurements of nJC–H and nJH–H in the aforesaid spin system, we have developed a new technique, named BASHD‐J‐resolved‐HMBC. This method incorporates band selective homo decoupled pulse and J‐scaling pulse into HMBC. In this method, high resolution cross peaks can be observed along the F1 axis by J‐scaling pulse, and band selective homo decoupled pulse simplified multiplet signals. Determinations of nJC–H and nJH–H of multiplet signals can easily be performed using the proposed pulse sequence. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

3.
Proton–proton 3J, 4J and 5J NMR coupling constants have been calculated for cyclohexane and monosubstituted cyclohexane conformers (substitiuents: Li, CH3, OH, F) by the two methods mentioned. Comparing the two methods on the basis of group theory, we show the necessity to use the second. The results from this method are compared with those of the literature.  相似文献   

4.
The 13C NMR spectra of pure exo-2-norbornyltrimethylstannane and a mixture of the exo- and endo-isomers have been recorded. 1H–13C polarization transfer spectra have been obtained and require the previously reported assignments for C-3 and C-4 in the exo-isomer to be reversed. The reported assignments for the endo-isomer are correct. The new assignment for C-4-exo [with J(119Sn,13C) vic=12 Hz, instead of the previously assigned J(vic)=23 Hz], has a very minor effect on the nature of the Karplus curve [for 3J(119Sn,13C)] generated previously.  相似文献   

5.
d ‐Glucaric acid (GA) is an aldaric acid and consists of an asymmetric acyclic sugar backbone with a carboxyl group positioned at either end of its structure (i.e., the C1 and C6 positions). The purpose of this study was to conduct a conformation analysis of flexible GA as a solution in deuterium oxide by NMR spectroscopy, based on J‐resolved conformation analysis using proton–proton (3JHH) and proton–carbon (2JCH and 3JCH) coupling constants, as well as nuclear overhauser effect spectroscopy (NOESY). The 2JCH and 3JCH coupling constants were measured using the J‐resolved heteronuclear multiple bond correlation (HMBC) NMR technique. NOESY correlation experiments indicated that H2 and H5 were in close proximity, despite the fact that these protons were separated by too large distance in the fully extended form of the chain structure to provide a NOESY correlation. The validities of the three possible conformers along the three different bonds (i.e., C2? C3, C3? C4, and C4? C5) were evaluated sequentially based on the J‐coupling values and the NOESY correlations. The results of these analyses suggested that there were three dominant conformers of GA, including conformer 1 , which was H2H3:gauche, H3H4:anti, and H4H5:gauche; conformer 2 , which was H2H3:gauche, H3H4:anti, and H4H5:anti; and conformer 3 , which was H2H3:gauche, H3H4: gauche, and H4H5:anti. These results also suggested that all three of these conformers exist in equilibrium with each other. Lastly, the results of the current study suggested that the conformational structures of GA in solution were ‘bent’ rather than being fully extended. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

6.
Improved pulse sequences for measuring long‐range C‐H coupling constants (nJC‐H), named selective COSY‐J‐resolved HMBC‐1 and ?2, have been developed. In the spin systems, such as ‐CHC‐CHA(CH3)‐CHB‐, a methine proton HA splits into a multiplet owing to several vicinal couplings with protons, resulting in attenuation of its cross‐peak intensity. Therefore, the measurements of nJC‐H with HA are generally difficult in the J‐resolved HMBC or selective J‐resolved HMBC spectrum. With the aim of accurate measurements of nJC‐H in such a spin system, we have developed new pulse sequences, which transfer the magnetization of a methyl group to its adjacent methine proton. The proposed pulse sequences successfully enable to enhance the sensitivity of HA cross peak in comparison with the selective J‐resolved HMBC pulse sequence. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

7.
Molecular interactions between uracil and nitrous acid (U–NA) [C4N2O2H4? NO2H] have been studied using B3LYP, B3PW91, and MP2 methods with different basis sets. The optimized geometries, harmonic vibrational frequencies, charge transfer, topological properties of electron density, nucleus‐independent chemical shift (NICS), and nuclear magnetic resonance one‐ and two‐bonds spin–spin coupling constants were calculated for U–NA complexes. In interaction between U and NA, eight cyclic complexes were obtained with two intermolecular hydrogen bonds N(C)HU…N(O) and OHNA…OU. In these complexes, uracil (U) simultaneously acts as proton acceptor and proton donor. The most stable complexes labeled, UNA1 and UNA2, are formed via NH bond of U with highest acidity and CO group of U with lowest proton affinity. There is a relationship between hydrogen bond distances and the corresponding frequency shifts. The solvent effect on complexes stability was examined using B3LYP method with the aug‐cc‐pVDZ basis set by applying the polarizable continuum model (PCM). The binding energies in the gas phase have also been compared with solvation energies computed using the PCM. Natural bond orbital analysis shows that in all complexes, the charge transfer takes place from U to NA. The results predict that the Lone Pair (LP)(O)U → σ*(O? H) and LP(N(O)NA → σ*(N(C)? H)U donor–acceptor interactions are most important interactions in these complexes. Atom in molecule analysis confirms that hydrogen bond contacts are electrostatic in nature and covalent nature of proton donor groups decreases upon complexation. The relationship between spin–spin coupling constant (1hJHY and 2hJHY) with interaction energy and electronic density at corresponding hydrogen bond critical points and H‐bonds distances are investigated. NICS used for indicating of aromaticity of U ring upon complexation. © 2013 Wiley Periodicals, Inc.  相似文献   

8.
The reaction of [RuCl2(PPh3)3] and closo-[B10H10]2? with p-IPhCOOH in CH2Cl2 solution affords two para-iodobenzoate exo-cyclized 11-vertex closo-ruthenaborane clusters [(PPh3)(p-IPhCO2)2RuB10H8] (1) and [(PPh3)2ClRu(PPh3)(p-IPhCO2)RuB10H9]?···?CH2Cl2 (2) that have been characterized by elemental analysis, FT-IR, 1H and 13C?NMR spectra and single-crystal X-ray diffraction analysis. Both clusters are based on a closo-type C 2 v 1?:?2?:?4?:?2?:?2 RuB10 stack with the metal occupying the unique six-connected apical position. In 1, the metal center has three exo-polyhedral ligands: one triphenylphosphine and two native oxygen atoms of para-iodobenzoates. The other oxygen atoms of two para-iodobenzoates are additionally bonded to B(2) and B(3) atoms respectively, resulting in two exo-cyclic five-membered Ru–O–C–O–B rings and engendering a symmetrical conformation. For 2, the metal center also has three exo-polyhedral ligands, one triphenylphosphine and one para-iodobenzoate to form one exo-cyclic five-membered Ru–O–C–O–B ring. There is an additional exo-polyhedral ruthenium atom bonding to the {RuB10} center via a {Ru–Ru} linkage and two {RuH μ B} bridges resulting in one closo distorted exo-polyhedral Ru(1)–Ru(2)–B(2)–B(4) tetrahedron.  相似文献   

9.
Two complexes [Ln(e,a-cis-1,4-chdc)(e,a-cis-1,4-Hchdc)(phen)(H2O)]2?10H2O (Ln = Eu, 1; Tb, 2, 1,4-H2chdc = 1,4-cyclohexanedicarboxylic acid; phen = 1,10-phenanthroline) have been synthesized and structurally characterized by single-crystal X-ray diffraction. Both complexes are doubly e,a-cis-1,4-chdc-bridged dimers. The e,a-cis-1,4-Hchdc, phen, and water molecules bond to Ln3+, forming nine-coordinate complexes. 3-D supramolecular frameworks are constructed by hydrogen bonds and π–π stacking interactions. Luminescence spectra exhibit the 5D07F J (J = 0–4) and 5D47F J (J = 6–3) transitions of Eu3+ for 1 and Tb3+ ion for 2, respectively.  相似文献   

10.
The comparative behaviour of the endo- and exo-norborneols and diastereomeric derivatives (acetates and benzoates) towards the NH3/NH4+ system was investigated. It appears that the proton affinity (PA) of the substrate relative to Pa(NH3) strongly influences competition between the protonation and nucleophilic substitution processes yielding the MH+ and [M + NH4 ? H2O]+ ions, respectively. Tandem mass spectrometry was used to compare collision-activated dissociation spectra of [M + NH4 ? H2O]+ with those of analogous endo- and exo-norbornylamines protonated in the source. This demonstrates that an SNimechanism occurs specifically for the isomeric norborneols; in contrast, for acetates and benzoates, stereospecific SNi and SN2 pathways take place for exo and endo derivatives, respectively. This particular behaviour is explained by considering the steric effect induced by the endo-H at C(6). In addition, the competitive decompositions of [M + NH4 – H2O]+ into NH4+ and [C7H11]+ daughter ions are consistent with the formation of a proton-bound complex intermediate. The observed stereochemical effects for these dauther ions are rationalized by means of arguments based on the estimated heats of formation of the transition states, which is lower for the exo-norbonyl protonated amine, consistent with anchimeric assistance, rather than a stepwise pathway which is proposed for the endoisomer.  相似文献   

11.
Vicinal C,H spin coupling (3JC,H) in substituted alkenes has been investigated systematically. Emphasis is laid on the stereochemical significance (Jtrans/Jcis) and on the various structural factors which influence 3JC,H, such as π-bond order, torsional angle ?, bond angle θ, electronegativity of substituents and steric effects. A new type of γ-effect is observed in 3JtransC,H which appears to have the same origin as the γ-shift effect. By comparison of 3J and 3JH,H, it was found that the relation 3JC,H ≈? 0·6 3JH,H holds for both trans and cis coupling constants. Finally, it is concluded that 3JC,H constitutes a valuable criterion to distinguish E- and Z-isomers, particularly in trisubstituted alkenes. Applications to natural products are presented.  相似文献   

12.
Contributions to the Chemistry of Phosphorus. 170. Constitutional and Configurational Isomers of Hexaphosphane(8), P6H8 Phosphane mixtures containing 5–10 P-% of hexaphosphane(8), P6H8, are obtained by thermolysis of diphosphane, P2H4, or as residue from distillation of crude diphosphane [3]. By complete analysis of the 31P{1H}-NMR spectrum on the basis of selective population transfer experiments, the following P6H8-isomers with a branched phosphorus skeleton have been identified and structurally characterized: the two diastereomers of 2-phosphinopentaphosphane ( 1a : erythro; 1b : threo), two of the three diastereomers of 3-phosphinopentaphosphane ( 2a : erythro, erythro; 2b : erythro, threo), and the highly symmetric 2,3-diphosphinotetraphosphane ( 3 ). The correlation between the diastereomers and the observed spin systems results from the preferred gauche orientation of neighboring free electron pairs, the dependence of 1J(PP) on dihedral angles as well as the 3J(PP) and 4J(PP) long range couplings. Any indications of the diastereomers of n-P6H8 with an unbranched chain of phosphorus atoms have not been found.  相似文献   

13.
Preparation of unsaturated sugars phosphonates using nucleophilic conjugate addition Different types of phosphorus nucleophiles underwent conjugate addition reaction with one of the branched-chain sugars 4, 5 or 11 the addition taking place either on the endo or the exo face of the furanose ring (or on both faces in the case of 11 ). The configuration at C(3) of these new phosphorus-bearing types of sugars as well as the configuration at the phosphorus atom of the cyclic phosphinates 9 and 10 was established by NMR. (3JP,H–C(2), 3JP,C(1)). Small amounts (7%) of the spiro enol phosphonate 16 were formed when 11 reacted with trimethyl phosphite.  相似文献   

14.
At one extreme of the proton‐transfer spectrum in cocrystals, proton transfer is absent, whilst at the opposite extreme, in salts, the proton‐transfer process is complete. However, for acid–base pairs with a small ΔpKa (pKa of base ? pKa of acid), prediction of the extent of proton transfer is not possible as there is a continuum between the salt and cocrystal ends. In this context, we attempt to illustrate that in these systems, in addition to ΔpKa, the crystalline environment could change the extent of proton transfer. To this end, two compounds of salicylic acid (SaH) and adenine (Ad) have been prepared. Despite the same small ΔpKa value (≈1.2), different ionization states are found. Both crystals, namely adeninium salicylate monohydrate, C5H6N5+·C7H5O3?·H2O, I , and adeninium salicylate–adenine–salicylic acid–water (1/2/1/2), C5H6N5+·C7H5O3?·2C5H5N5·C7H6O3·2H2O, II , have been characterized by single‐crystal X‐ray diffraction, IR spectroscopy and elemental analysis (C, H and N) techniques. In addition, the intermolecular hydrogen‐bonding interactions of compounds I and II have been investigated and quantified in detail on the basis of Hirshfeld surface analysis and fingerprint plots. Throughout the study, we use crystal engineering, which is based on modifications of the intermolecular interactions, thus offering a more comprehensive screening of the salt–cocrystal continuum in comparison with pure pKa analysis.  相似文献   

15.
A tabulation has been compiled for twenty 13C? H coupling constants of various carboxylic acids and includes 2J(C,H), 3J(C,H) and 4J(C,H) values of olefins (both cis and trans); 3J(C,H), 4J(C,H) and 5J(C,H) values of aromatics; 3J(C,H) and 4J(C,H) values of acetylenes; and 2J(C,H) and 3J(C,H) values of rigid aliphatics. This tabulation has been completed in the present study by the spin-tickling proton n.m.r. study of 13C-carboxyl-endo-1,2,3,4,7,7-hexachloronorbornene-5-carboxylic acid, which has established that the 2J(C,H) value is negative and the 3J(C,H) values (both cis and trans) are positive in this system. A plot of these twenty J(C,H) values vs the corresponding J(H,H) values of geometrically equivalent model systems (where there is a proton in place of a carboxyl group) gives a correlation coefficient of 0·975 (with a slope of 0·62), indicating that carbon–proton and proton–proton couplings operate by similar mechanisms throughout this broad series of structural types.  相似文献   

16.
Long‐range coupling constants 5JHortho,OMe were measured in series of methoxyindoles, methoxycoumarins, and methoxyflavones by the modified J doubling in the frequency domain method. The COSY and NOESY spectra revealed the coupling of the –OMe group with a specific proton at the ortho position and its preferred conformation. Homonuclear 1H–1H couplings were confirmed by irradiation of the –OMe signal. Density functional theory calculations of 5JHortho,OMe using the modified aug‐cc‐pVTZ basis set evidenced that the Fermi contact term shows good agreement with the experimental J values. Accurate chemical shift and coupling constant values followed after iterative quantum mechanical spectral analysis using the PERCH software. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

17.
13C, 1H spin coupling constants of dimethylacetylene have been determined by the complete analysis of the proton coupled 13C NMR spectrum. For the methyl carbon 1J(CH) = + 130.64 Hz and 4J(CH) = + 1.58 Hz, and for the acetylenic carbon 2J(CH) = ? 10.34 Hz and 3J(CH) = +4.30 Hz. The 5J(HH) long-range coupling constant (+2.79 Hz) between the methyl protons was also determined.  相似文献   

18.
High-resolution proton spectra of acridine and the five monomethyl derivatives have been analyzed using multiple resonance and computer techniques. It was possible to determine the magnitudes and relative signs of the inter-ring coupling constants between H-C(9) and all the protons of the adjacent rings. Positive signs have been obtained for 5J29 and 5J49, negative signs for 4J19 and 6J39. The coupling constants are discussed in terms of σ- and π-contributions and as a function of the geometry of the pathway.  相似文献   

19.
The difficult Diels-Alder additions of α-acetoxy- and α-chloroacrylonitrile to furan can be run at 20–35° and atmospheric pressure in the presence of CuCl. Cu(BF4) · 6 H2O, Cu(OOCCH3)2 · H2O or cupric tartrate · 3H2O. Under kinetic control, the exo-carbonitrile adducts 2 and 8 , respectively, are favoured. Saponification of the 2endo-acetoxy-7-oxabicyclo[2.2.1]hept-5-ene-2exo-carbonitrile ( 2 ) furnished the 7-oxabicyclo[2.2.1]hept-5-en-2-one ( 4 ). Basic hydrolysis of the adducts ( 8 + 9 ) of α-chloroacrylonitrile to furan and its 5exo, 6exo-isopropylidenedioxy derivatives did not give the corresponding ketones, the carboxamides 14 + 15 and 16 + 17 , respectively, were isolated.  相似文献   

20.
The properties of the intramolecular hydrogen bonds of doubly 15N‐labeled protonated sponges of the 1,8‐bis(dimethylamino)naphthalene (DMANH+) type have been studied as a function of the solvent, counteranion, and temperature using low‐temperature NMR spectroscopy. Information about the hydrogen‐bond symmetries was obtained by the analysis of the chemical shifts δH and δN and the scalar coupling constants J(N,N), J(N,H), J(H,N) of the 15NH15N hydrogen bonds. Whereas the individual couplings J(N,H) and J(H,N) were averaged by a fast intramolecular proton tautomerism between two forms, it is shown that the sum |J(N,H)+J(H,N)| generally represents a measure of the hydrogen‐bond strength in a similar way to δH and J(N,N). The NMR spectroscopic parameters of DMANH+ and of 4‐nitro‐DMANH+ are independent of the anion in the case of CD3CN, which indicates ion‐pair dissociation in this solvent. By contrast, studies using CD2Cl2, [D8]toluene as well as the freon mixture CDF3/CDF2Cl, which is liquid down to 100 K, revealed an influence of temperature and of the counteranions. Whereas a small counteranion such as trifluoroacetate perturbed the hydrogen bond, the large noncoordinating anion tetrakis[3,5‐bis(trifluoromethyl)phenyl]borate B[{C6H3(CF3)2}4]? (BARF?), which exhibits a delocalized charge, made the hydrogen bond more symmetric. Lowering the temperature led to a similar symmetrization, an effect that is discussed in terms of solvent ordering at low temperature and differential solvent order/disorder at high temperatures. By contrast, toluene molecules that are ordered around the cation led to typical high‐field shifts of the hydrogen‐bonded proton as well as of those bound to carbon, an effect that is absent in the case of neutral NHN chelates.  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