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.
All J(P? H) and J(P? C) values, including signs, have been obtained in acetylenic and propynylic phosphorus derivatives, R2P(X)? C?C? H and R2P(X)? C?C? CH3 (X ? oxygen, lone pair and R ? C6H5, N(CH3)2, OC2H5, N(C6H5)2, Cl) from 1H and 13C NMR spectra. In PIV derivatives the following signs are obtained: 1J(P? C)+, 2J(P? C)+, 3J(P? C)+, 3J(P? H)+, 4J(P? H)? . Linear relations are observed between 1J(P? C), 2J(P? C) and 3J(P? C) versus 3J(P? H), indicating that these coupling constants are mainly dependent on the Fermi contact term, though the other terms of the Ramsey theory do not seem to be negligible for 1J(P? C) and 2J(P? C). In PIII derivatives these signs are: 1J(P? C)- and +, 2J(P? C)+, 3J(P? C)-, 3J(P? H)-, 4J(P? H)+. Only 3J(P? C) and 3J(P? H) reflect a small contribution of the Fermi contact term while in 1J(P? C) and 2J(P? C) this contribution seems to be negligible relative to the orbital and/or spin dipolar coupling mechanisms.  相似文献   

2.
Synthesis and NMR Spectra of Some 13C-Labelled Thio- and Seleno-ethers, -acetals, and -orthoesters Twenty-seven different open-chain and cyclic derivatives (RX)nCH4-n and (RX)nCH3-nR′ with n = 1?3, X = S or Se, R,R′ = alkyl or aryl, 1,3,5-trithiane, and bis-(dimethylsulfonio)methane and -methanide with single or multiple 13C-labelling have been synthesized. The 13C-NMR spectra of the sulfur and selenim compounds have been measured, and the dependence of the chemical shifts (δc) and coupling constants [′J(C,H), ′J(Se,C)] from the substitution pattern in discussed (Fig. 1) and compared with the polyhalogeno-methanes (Fig. 2).  相似文献   

3.
J(13C1H) coupling constants for some methyl- and aminopyrimidines have been determined by 13C NMR. Both the one-bond and long-bond and long-range coupling constants follow general trends which can be summarized in a few simple rules. In particular, the 3J(C-i,H) coupling constants between a ring carbon C-i and the ring protons are larger than the 2J(C-i,H) coupling constants. The opposite is observed for the couplings between the ring carbons and the methyl protons: 3J(C,Me). These general rules are very useful for the assignment of resonances in complex 13C spectra of pyrimidines and seem to be valid for other 6-membered aromatic nitrogen heterocycles. Furthermore, the additivity of substituent effects on 1J (CH) for monosubstituted pyrimidines allows the estimation of 1J (CH) for polysubstituted pyrimidines with a very good accuracy.  相似文献   

4.
Substituent Effects on NMR Spectra of Pentafulvenes. 13C, 13C-NMR Coupling Constants (1J(C, C)) 1H- and 13C-NMR spectra of 6-monosubstituted pentafulvenes 1 – 8 have been analysed, and 1J(C, C) coupling constants have been determined from ID-inadequate spectra of 13C satellites. It turns out that 13C,13C coupling constants of the ring C-atoms, and especially J(1,2)/J(3,4) and J(2,3), reflect the extent of π delocalisation in the fulvene ring. With increasing electron-donating capacity of the substituent R, J(1,2)/J(3,4) values are decreasing, while J(2,3) (and J(1,5)/J(4,5) as well) are increasing, and linear correlations of Hammett substituent constants σ+ and 1J(C,C) values are obtained.  相似文献   

5.
1J(13C?13C) nuclear spin–spin coupling constants in derivatives of acetylene have been measured from natural abundance 13C NMR spectra and in one case (triethylsilyllithiumacetylene) from the 13C NMR spectrum of a 13C-enriched sample. It has been found that the magnitude of J(C?C) depends on the electronegativity of the substituents at the triple bond. The equation 1J(13C?13C) = 43.38 Ex + 17.33 has been derived for one particular series of the compounds Alk3SiC?CX, where X denotes Li, R3Sn, R3Si, R3C, I, Br or Cl. The 1J(C?C) values found in this work cover a range from 56.8 Hz (in Et3SiC?Li) to 216.0 Hz (in PhC?CCI). However, the 1J(C?C) vs Ex equation combined with the Egli–von Philipsborn relationship allows the calculation of the coupling constants in Li2C2 (32 Hz) and in F2C2 (356 Hz). These are probably the lowest and the highest values, respectively, which can be attained for 1J(CC) across a triple bond. The unusually large changes of the 1J(C?C) values are explained in terms of substituent effects followed by a re-hybridization of the carbons involved in the triple bond. INDO FPT calculations performed for two series of acetylene derivatives, with substituents varied along the first row of the Periodic Table, corroborate the conclusions drawn from the experimental data.  相似文献   

6.
The 1H and 31P NMR spectra of a series of compounds containing the PIII-N-PV skeleton including Cl2PNMeP(Z)Cl2 (Z?O or S), Cl2P·NMe·P(O)CINMe2, P4(NMe)6Zn (Z?S, n = 1?3; Z?Se, n = 1), XP(NBut)2P(Z)X (X?Cl, Z?O or S; X?OMe, Z?S or Se; X?NMe2, Z?S or Se; X?NEt2, Z?Se), (X?O or S), and Ph2PNRP(S)Ph2 (R?Me, Et) have been obtained. 1H-{31P} double resonance, and in selected cases, 31P-{1H, 77Se} and 31P-{1H, 31P} triple resonance experiments, indicate that 2J(PIII NPV) is positive in acyclic compounds, negative in most cyclic or cage compounds, and furthermore, is related to the conformation adopted by the PIII-N bond.  相似文献   

7.
The proton and carbon-13 NMR spectra of thirteen trialkylmetal derivatives of pyridine, several of which were previously unknown, have been recorded and analysed. The proton NMR spectra show variations in proton chemical shifts but not in proton-proton coupling constants when the metal substituent is changed; the ring proton-metal coupling constants nJ(M? H) in the tin and lead derivatives correspond closely with the corresponding proton-proton couplings nJ(H? H) in pyridine. The carbon-13 chemical shifts of the carbons bound to the metal can apparently be correlated with the electron-donating ability of the trialkylmetal group. In the trimethylstannylpyridines the value of 1J(Sn? Cring) varies greatly with the position of the Me3Sn group.  相似文献   

8.
Photolysis of a benzene solution containing [Fe3(CO)93-E)2] (E=S, Se), [(η5-C5R5)Fe(CO)2(CCRI)] (R=H, Me; RI=Ph, Fc), H2O and Et3N results in formation of new metal clusters [(η5-C5R5)Fe3(CO)63-E)(μ3-ECCH2RI)] (R=H, RI=Ph, E=S 1 or Se 2; R=Me, RI=Ph, E=S 3 or Se 4; R=H, RI=Fc, E=S 5; R=Me, RI=Fc, E=S 6 or Se 7). Reaction of [Fe3(CO)93-S)2]with [(η5-C5R5)Mo(CO)3(CCPh)] (R=H, Me), under same conditions, produces mixed-metal clusters [(η5-C5R5)MoFe2(CO)63-S)(μ-SCCH2Ph)] (R=H 8; R=Me 9). Compounds 19 have been characterised by IR and 1H and 13C-NMR spectroscopy. Structures of 1, 5 and 9 have been established crystallographically. A common feature in all these products is the formation of new C-chalcogen bond to give rise to a (ECCH2RI) ligand.  相似文献   

9.
The influence of exocyclic substituents on π‐delocalization of pentafulvenes 2 , heptafulvenes 3 , and nonafulvenes 4 has been investigated. Pentafulvenes 2 : Changes of bond lengths (induced by exocyclic substituents R1 and R2 of 2 ) are reflected by systematic changes of 3J(H,H) (Fig. 2) as well as of 1J(C,C) coupling constants (Fig. 4), so that linear correlations of σp+ vs. 3J(H,H) and 1J(C,C) coupling constants were obtained. Plots of that type are very useful for determining the extent of π‐delocalization of various pentafulvalenes 5 – 8 (Figs. 6 and 12). Charge density effects of pentafulvenes and pentafulvalenes were observed by substituent‐induced shifts of the ring C‐atoms (Fig. 5). Heptafulvenes 3 : Contrary to planar pentafulvenes, heptafulvenes did not show any linear correlations of σp+ vs. 3J(H,H)‐plots (Fig. 8) or σp+ vs. δ(13C)‐plots (Fig. 9), although substituents R1, R2 clearly influenced 3J(H,H)‐coupling constants as well as 13C chemical shifts of the ring H‐atoms and ring C‐atoms. In the NMR spectra of ‘heptafulvenes with inverse ring polarization’ (in the lower range of Fig. 8), 3J(H,H)‐coupling constants were strongly alternating and were barely influenced by exocyclic substituents. This supported a boat conformation of the corresponding heptafulvenes. In the range of Hammett σp+values above ?0.5 to 0, strong substituent effects started to be effective, and a nearly linear approach of 3J(H,H)‐coupling constants J(2,3)/J(4,5) and J(3,4) was observed. This meant that, as soon as heptafulvenes were planar or nearly planar, there existed similar substituent effects as for planar pentafulvenes. – A similar ‘turning point’ was observed in plots of σp+ vs. 13C‐chemical shifts around σp+=0 (Fig. 9): In the range of strong electron‐accepting groups (above σp+=1), there was a marked substituent‐induced high‐frequency shift which strongly decreased in the series C(7)>C(2)/C(5)>C(3)/C(4), while C(1)/C(6) was barely influenced. Nonafulvenes 4 : Most nonafulvenes are non‐planar olefins with strongly alternating vicinal H,H‐coupling constants. This has been convincingly shown by the high‐resolution 1H‐NMR spectrum of 10‐dimethylaminononafulvene ( 4c , Fig. 10), which was not planar but contained a nearly planar (E)dienamine substructure of the segment C(7)?C(8)? C(9)?C(10)? NMe2 according to the NMR data. Only with very strong π‐donors (like two dimethylamino groups in 4b ), planarization of the nine‐membered ring could be observed at low temperatures (Fig. 10). Finally, the first stable nonatriafulvalene (11,12‐bis(diethylamino)nonatriafulvalene ( 10 )) existed in the planar dipolar form in the whole temperature range and even in unpolar solvents.  相似文献   

10.
The variation in the one–bond couplings 1J(CH) in vinyl derivatives with substituent has been examined. For the geminal proton 1J correlates very badly with substituent electronegativity but extremely well with σI, if conjugating substituents are excluded. In the case of halogen substituents the marked stereospecificity of 1J(CH) for the cis and trans protons can be rationalised in terms of an intrinsic dependence of πCH on the dihedral angle between the coupling atoms and the perturbing substituent, with an additional positive increment to the cis coupling due to direct interaction of the substituent non-bonding electrons or to orbital circulation of the substituent electrons. The intrinsic specificity of β-substituent effects on 1J(CH) is also found in analogous compounds containing C?N and C?O bonds.  相似文献   

11.
The 1H and 13C NMR spectra of 1,2-dibromoethane-13C2 have been analyzed to determine the magnitude (38·9 Hz) and sign (positive) of 1J(C? C) relative to those of 3J(H? H) (positive). This type of coupling appears to be rather insensitive to the presence of bromine or methyl as substituents on the carbons.  相似文献   

12.
Alkynyl mercury compounds of the type Hg(CCR)2 (I), R′HgCCR (II) and R′HgCCHgR′ (III) have been studied by 1H, 13C and 199Hg NMR. Chemical shifts (δ1H, δ13C, δ199Hg) and coupling constants J(199Hg1H), J(199Hg13C), J(13C1H) and J(13C13C) (in natural abundance) are reported. The changes in magnitude of the coupling constants 1J(199Hg13C) and 1J(13C13C) cannot be fully explained in terms of changes in the “s-character” of the HgC bond and the CC bond, respectively. The shielding of the alkynyl carbon linked to mercury in II is decreased by ca. 23 ppm as compared to the analogous carbon in I. This indicates a greatly different degree of polarization for the HgC bonds in I and II in agreement with the behaviour of 1J(199Hg13C) and 1J(13C13C). The solvent and temperature dependence of the 199Hg chemical shift of I (R = C6H5, C4H9n) and II (R = H, R′= CH3) has been studied. The results indicate covalent interactions of I with amines, pyridine, dimethylsulphide and Br, while the interaction with acetonitrile and oxygen donors (DMSO, DMF, dioxane, acetone) is of a different nature.  相似文献   

13.
Experimental and theoretical studies on equilibria between iridium hydride alkylidene structures, [(TpMe2)Ir(H){?C(CH2R)ArO }] (TpMe2=hydrotris(3,5‐dimethylpyrazolyl)borate; R=H, Me; Ar=substituted C6H4 group), and their corresponding hydride olefin isomers, [(TpMe2)Ir(H){R(H)C? C(H)OAr}], have been carried out. Compounds of these types are obtained either by reaction of the unsaturated fragment [(TpMe2)Ir(C6H5)2] with o‐C6H4(OH)CH2R, or with the substituted anisoles 2,6‐Me2C6H3OMe, 2,4,6‐Me3C6H2OMe, and 4‐Br‐2,6‐Me2C6H2OMe. The reactions with the substituted anisoles require not only multiple C? H bond activation but also cleavage of the Me? OAr bond and the reversible formation of a C? C bond (as revealed by 13C labeling studies). Equilibria between the two tautomeric structures of these complexes were achieved by prolonged heating at temperatures between 100 and 140 °C, with interconversion of isomeric complexes requiring inversion of the metal configuration, as well as the expected migratory insertion and hydrogen‐elimination reactions. This proposal is supported by a detailed computational exploration of the mechanism at the quantum mechanics (QM) level in the real system. For all compounds investigated, the equilibria favor the alkylidene structure over the olefinic isomer by a factor of between approximately 1 and 25. Calculations demonstrate that the main reason for this preference is the strong Ir–carbene interactions in the carbene isomers, rather than steric destabilization of the olefinic tautomers.  相似文献   

14.
Infinite dilution 29Si and 13C NMR chemical shifts were determined from concentration dependencies of the shifts in dilute chloroform and acetone solutions of para substituted O‐silylated phenols, 4‐R‐C6H4‐O‐SiR′2R″ (R = Me, MeO, H, F, Cl, NMe2, NH2, and CF3), where the silyl part included groups of different sizes: dimethylsilyl (R′ = Me, R″ = H), trimethylsilyl (R′ = R″ = Me), tert‐butyldimethylsilyl (R′ = Me, R″ = CMe3), and tert‐butyldiphenylsilyl (R′ = C6H5, R″ = CMe3). Dependencies of silicon and C‐1 carbon chemical shifts on Hammett substituent constants are discussed. It is shown that the substituent sensitivity of these chemical shifts is reduced by association with chloroform, the reduction being proportional to the solvent accessible surface of the oxygen atom in the Si‐O‐C link. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

15.
The measurement of the magnitude and sign of 2J(C,H) couplings offers a reliable way to determine the absolute configuration at a carbon center in a fixed cyclic system. A decrease of the dihedral angle ? in the O—CA—CB—H fragment always leads to a change of the 2J(CA,HB) coupling to more negative values, independent of the type and position of substituents at the two carbon centers. The orientations of the two substituents at C‐3 of the epimeric pair 1 and 2 were determined unambiguously through the measurement of the geminal coupling constants between C‐3 and the hydrogen atoms at C‐2 and C‐4. In particular, 2J(C‐3,H‐2ax) with ?1.5 Hz, ? = 174° in 1 and ?6.6 Hz, ? = 47° in 2 , and 2J(C‐3,H‐4) with +1.5 Hz, ? = 175° in 1 and ?4.7 Hz, ? = 49° in 2 showed the greatest differences between the two epimers. Both couplings therefore allow the determination of the absolute configuration at C‐3. It should be noted, however, that the size of the coupling constants can be different for dihedral angles of nearly identical size, when there are different numbers of electronegative substituents on the two coupling pathways, i.e. no O‐substituent at C‐2, but one axial O‐substituent at C‐4. It becomes clear that it is not sufficient to measure the magnitude of 2J coupling constants only, but that the sign of the geminal coupling is needed to identify the absolute configuration at a chiral center. The coupling of C‐3 with H‐2eq is not useful for the determination of the configuration at C‐3, as the similarity of the dihedral angles ? (O—C‐3—C‐2—H‐2eq) (57° in 1 and 70° in 2 ) leads to identical coupling constants (?6.1 Hz) for both epimers. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

16.
1J(15N13C) values obtained from FT 13C NMR spectra were measured for a number of 15N-enriched aniline derivatives and are found to exhibit varying degrees of dependence on the nature of the ring substituent. Theoretical calculations of 1J(15N13C) values for representative members of the systems examined were made using INDO parameters and a ‘sum-over-states’ perturbation approach. The calculated coupling constants are generally in fair agreement with experimental values when the integral products SN2(o)SC2(o) and (r?3)N(r?3)C have values of 34.437 au?6 and 2.770 au?6, respectively.  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.
1H- and 13C-NMR spectra of a series of 8-R1-substituted as well as of 8,8-R1, R2-disubstituted heptafulvenes, varying from inversely polarized ( 3l ) to unpolar ( 3h ) and polar heptafulvenes with electron-withdrawing groups ( 3d , e , f ), have been analyzed and compared with those of methoxytropylium salt 5a . The results concerning 3J (H,H) values and 13C-chemical shifts are shown in Figs. 5 and 6. It turns out that all the NMR parameters are strongly influenced by substituents R1, R2, but contrary to planar pentafulvenes, no linear correlations of the NMR parameter vs. Hammett substituent constant σ+ are obtained in the series 3l → 3d . 3J coupling constants J(2,3)/J(4,5) and J(3,4) are not much influenced by substituent changes in the series 3l → 3h , but are approaching in the row 3h → 3d . Similarly, signals of the 13C-atoms undergo a moderate shift to higher frequencies in the row 3l → 3h , but are strongly influenced by ? M groups, whereby the sensitivity is decreasing in the series C(7) > C(2)/C(5) > C(3)/C(4) > C(1)/C(6). These results are essentially explained by a boat conformation of inversely polarized heptafulvenes of the type 3l and an increasing planarization of the ring on going to polar heptafulvenes of type 3d .  相似文献   

19.
The spin-spin coupling constants J(H? H) and J(Se? H) of 2- and 3-substituted selenophenes, whose signs have been obtained by double resonance experiments, have been correlated with the reactivity constants F and R of Swain and Lupton by means of linear equations J = i + fF + rR. The relative inductive and mesomeric contributions to the coupling constants are discussed. Substituent effects on J(H? H) and J(Se? H) are found to be additive in 2,4-disubstituted selenophenes. In agreement with experimental results, this additivity relationship indicates that 3J(Se? H) becomes negative in 2,4-dinitroselenophene. Evidence is given from long range coupling constant data that 2-formylselenophene exists almost exclusively in the Se? O cis conformation and 3-formylselenophene in the Se? O trans conformation.  相似文献   

20.
From a carbon magnetic resonance study of several alkylcobaloximes RCo(DMG)2B (DMG = dimethylglyoximate monoanion), it was possible to estimate the α, β and γ effects of the Co(DMG)2B group on the chemical shifts of the carbon atoms of various alkyl groups R. The chemical shifts of the carbon atoms belonging to the equatorial ligands and to the axial base B are not significantly affected by structural modification of the R groups. Values of δ in benzylcobaloximes XC6H4CH2Co(DMG)2B agree with a donor effect of the ? CH2Co(DMG)2B radical. Values of 1J(13C? H) coupling constants, measured in 13C enriched methylcobaloximes, do not vary appreciably when B is changed (J(13C? H) = 137 ± 1 Hz) and are close to the value obtained for methylcobalamine.  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