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 874 毫秒
1.
This work reports on the comprehensive calculation of the NMR one‐bond spin–spin coupling constants (SSCCs) involving carbon and tellurium, 1J(125Te,13C), in four representative compounds: Te(CH3)2, Te(CF3)2, Te(C?CH)2, and tellurophene. A high‐level computational treatment of 1J(125Te,13C) included calculations at the SOPPA level taking into account relativistic effects evaluated at the 4‐component RPA and DFT levels of theory, vibrational corrections, and solvent effects. The consistency of different computational approaches including the level of theory of the geometry optimization of tellurium‐containing compounds, basis sets, and methods used for obtainig spin–spin coupling values have also been discussed in view of reproducing the experimental values of the tellurium–carbon SSCCs. Relativistic corrections were found to play a major role in the calculation of 1J(125Te,13C) reaching as much as almost 50% of the total value of 1J(125Te,13C) while relativistic geometrical effects are of minor importance. The vibrational and solvent corrections account for accordingly about 3–6% and 0–4% of the total value. It is shown that taking into account relativistic corrections, vibrational corrections and solvent effects at the DFT level essentially improves the agreement of the non‐relativistic theoretical SOPPA results with experiment. © 2016 Wiley Periodicals, Inc.  相似文献   

2.
Linear and nonlinear halogen dependencies of the 13C magnetic shielding constants of CH4−nIn, CH4−nBrn, CCl4−nIn, and CBr4−nIn were fairly reproduced by the ab initio generalized unrestricted Hartree–Fock (GUHF)/finite perturbation (FP) method including spin‐orbit (SO) interaction and spin‐free relativistic (SFR) terms. As seen from the experimental trends, the calculated 13C chemical shifts in CCl4−nIn and CBr4−nIn depend linearly on n=0–4, while those in CH4−nIn and CH4−nBrn depend nonlinearly. We found that both the linear and nonlinear dependencies are due to the relativistic effects, and especially due to the Fermi–Contact (FC) term originating from the SO interaction. © 2001 John Wiley & Sons, Inc. J Comput Chem 22: 528–536, 2001  相似文献   

3.
Lanthanum‐139 NMR spectra of stationary samples of several solid LaIII coordination compounds have been obtained at applied magnetic fields of 11.75 and 17.60 T. The breadth and shape of the 139La NMR spectra of the central transition are dominated by the interaction between the 139La nuclear quadrupole moment and the electric field gradient (EFG) at that nucleus; however, the influence of chemical‐shift anisotropy on the NMR spectra is non‐negligible for the majority of the compounds investigated. Analysis of the experimental NMR spectra reveals that the 139La quadrupolar coupling constants (CQ) range from 10.0 to 35.6 MHz, the spans of the chemical‐shift tensor (Ω) range from 50 to 260 ppm, and the isotropic chemical shifts (δiso) range from ?80 to 178 ppm. In general, there is a correlation between the magnitudes of CQ and Ω, and δiso is shown to depend on the La coordination number. Magnetic‐shielding tensors, calculated by using relativistic zeroth‐order regular approximation density functional theory (ZORA‐DFT) and incorporating scalar only or scalar plus spin–orbit relativistic effects, qualitatively reproduce the experimental chemical‐shift tensors. In general, the inclusion of spin–orbit coupling yields results that are in better agreement with those from the experiment. The magnetic‐shielding calculations and experimentally determined Euler angles can be used to predict the orientation of the chemical‐shift and EFG tensors in the molecular frame. This study demonstrates that solid‐state 139La NMR spectroscopy is a useful characterization method and can provide insight into the molecular structure of lanthanum coordination compounds.  相似文献   

4.
Calculations of nitrogen NMR parameters [chemical shifts δN and indirect nuclear spin–spin coupling constants J(N,N), J(N,13C), J(29Si,N)] of noncyclic azo‐compounds R1 NN R2 (R1, R2 = H, Me, Ph, SiH3, SiMe3) and cyclic azo‐compounds [NNCH2, NN(CH2)3 NN(CH2)2SiH2, and NN(SiH2CH2SiH2)] by density functional theory (DFT) methods [B3LYP/6‐311+G(d,p) level of theory] provide data in reasonable agreement with experimental values. The influence of cis‐ and trans‐geometry is reflected by the calculations, and amino‐nitrenes are also included for comparison. The spin–spin coupling constants are analyzed with respect to contact (Fermi contact term, FC) and non‐ contact contributions (paramagnetic and diamagnetic spin‐orbital terms, PSO and DSO, and spin‐dipole term, SD). Bis(trimethylsilyl)diazene 6a can be generated by an alternative method, using the reaction of bis(trimethylsilyl)sulfur diimide with bis‐ (trimethylsilyl)amino‐trimethylsilylimino‐phosphane. © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:84–91, 2005; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.20075  相似文献   

5.
The tellurium(II) dithiolates Te[SCH2CH2C(O)OCH3]2, ( 1 ), Te[SCH2CH2CH2SC(O)CH3]2, ( 2 ), and Te[SCH2CH2CH2CH2SC(O)CH3]2, ( 3 ) were synthesized from Te(StBu)2 and the corresponding thiol. All compounds are sensitive toward higher temperatures and light and decompose to elemental tellurium and the disulfide. In the solid state, the Te atom of 1 exhibits the novel Te(S2Te2) coordination mode. Additionally to the two Te—S bonds, each Te atom forms two long Te···Te contacts to neighboring molecules, leading to a coordination number of four and a distorted sawhorse configuration. No intramolecular Te···O interactions are present in the solid state, in accordance with ab initio calculations (MP2/ecp‐basis) for the isolated molecule. 125Te NMR shifts of all compounds lay within a narrow range and close to the respective shift of other Te(SCH2R)2 compounds. VT 125Te NMR spectra gave no hint to donor acceptor interactions in solution for any of the compounds and thus corroborate results from IR‐spectroscopy, ab initio geometry optimizations, and thermochemical calculations.  相似文献   

6.
The Te ⋅⋅⋅ Te secondary bonding interactions (SBIs) in solid cyclic telluroethers were explored by preparing and structurally characterizing a series of [Te(CH2)m]n (n=1–4; m=3–7) species. The SBIs in 1,7-Te2(CH2)10, 1,8-Te2(CH2)12, 1,5,9-Te3(CH2)9, 1,8,15-Te3(CH2)18, 1,7,13,19-Te4(CH2)20, 1,8,15,22-Te4(CH2)24 and 1,9,17,25-Te4(CH2)28 lead to tubular packing of the molecules, as has been observed previously for related thio- and selenoether rings. The nature of the intermolecular interactions was explored by solid-state PBE0-D3/pob-TZVP calculations involving periodic boundary conditions. The molecular packing in 1,7,13,19-Te4(CH2)20, 1,8,15,22-Te4(CH2)24 and 1,9,17,25-Te4(CH2)28 forms infinite shafts. The electron densities at bond critical points indicate a narrow range of Te ⋅⋅⋅ Te bond orders of 0.12–0.14. The formation of the shafts can be rationalized by frontier orbital overlap and charge transfer.  相似文献   

7.
The main factors affecting the accuracy and computational cost of the Second‐order Möller‐Plesset perturbation theory (MP2) calculation of 77Se NMR chemical shifts (methods and basis sets, relativistic corrections, and solvent effects) are addressed with a special emphasis on relativistic effects. For the latter, paramagnetic contribution (390–466 ppm) dominates over diamagnetic term (192–198 ppm) resulting in a total shielding relativistic correction of about 230–260 ppm (some 15% of the total values of selenium absolute shielding constants). Diamagnetic term is practically constant, while paramagnetic contribution spans over 70–80 ppm. In the 77Se NMR chemical shifts scale, relativistic corrections are about 20–30 ppm (some 5% of the total values of selenium chemical shifts). Solvent effects evaluated within the polarizable continuum solvation model are of the same order of magnitude as relativistic corrections (about 5%). For the practical calculations of 77Se NMR chemical shifts of the medium‐sized organoselenium compounds, the most efficient computational protocols employing relativistic Dyall's basis sets and taking into account relativistic and solvent corrections are suggested. The best result is characterized by a mean absolute error of 17 ppm for the span of 77Se NMR chemical shifts reaching 2500 ppm resulting in a mean absolute percentage error of 0.7%. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
Two‐component relativistic density functional theory (DFT) with the second‐order Douglas–Kroll–Hess (DKH2) one‐electron Hamiltonian was applied to the calculation of nuclear magnetic resonance (NMR) shielding constant. Large basis set dependence was observed in the shielding constant of Xe atom. The DKH2‐DFT‐calculated shielding constants of I and Xe in HI, I2, CuI, AgI, and XeF2 agree well with those obtained by the four‐component relativistic theory and experiments. The Au NMR shielding constant in AuF is extremely more positive than in AuCl, AuBr, and AuI, as reported recently. This extremely positive shielding constant arises from the much larger Fermi contact (FC) term of AuF than in others. Interestingly, the absolute values of the paramagnetic and the FC terms are considerably larger in CuF and AuF than in others. The large paramagnetic term of AuF arises from the large d‐components in the Au dπ –F pπ and Au sdσ–F pσ molecular orbitals (MOs). The large FC term in AuF arises from the small energy difference between the Au sdσ + F pσ and Au sdσ–F pσ MOs. The second‐order magnetically relativistic effect, which is the effect of DKH2 magnetic operator, is important even in CuF. This effect considerably improves the overestimation of the spin‐orbit effect calculated by the Breit–Pauli magnetic operator. © 2013 Wiley Periodicals, Inc.  相似文献   

9.
Interaction of the tetradentate redox-active 6,6′-[1,2-phenylenebis(azanediyl)]bis(2,4-di-tert-butylphenol) (H4L) with TeCl4 leads to neutral diamagnetic compound TeL ( 1 ) in high yield. The molecule of 1 has a nearly planar TeN2O2 fragment, which suggests the formulation of 1 as TeIIL2−, in agreement with the results of DFT calculations and QTAIM and NBO analyses. Reduction of 1 with one equivalent of [CoCp2] leads to quantitative formation of the paramagnetic salt [CoCp2]+[ 1 ].−, which was characterised by single-crystal XRD. The solution EPR spectrum of [CoCp2]+[ 1 ].− at room temperature features a quintet due to splitting on two equivalent 14N nuclei. Below 150 K it turns into a broad singlet line with two weak satellites due to the splitting on the 125Te nucleus. Two-component relativistic DFT calculations perfectly reproduce the a(14N) HFI constants and A(125Te) value responsible for the low-temperature satellite splitting. Calculations predict that the additional electron in 1 .− is localised mainly on L, while the spin density is delocalised over the whole molecule with significant localisation on the Te atom (≥30 %). All these data suggest that 1 .− can be regarded as the first example of a structurally characterised monomeric tellurium–nitrogen radical anion.  相似文献   

10.
2,2,6,6‐Tetramethylpiperidine‐1‐yloxyl derivatives substituted with either hydrogen bonding [‐OH, ‐OSO3H] or ionic [‐OSO3?Na+, ‐OSO3?K+, N+(CH3)3I?, N+(CH3)3 N?(SO2‐CF3)2] substituents are investigated in 1‐butyl‐3‐methylimidazolium tetrafluoroborate over a wide temperature range covering both glassy and viscous states. The Vogel–Fulcher–Tammann equation describes the temperature dependence of the ionic liquid viscosity. Quantum chemical calculations of the spin probes at the UB3LYP/6‐311(2d,p++) level are done to describe the dependence of the spin density on nitrogen on the substitution pattern of the 4‐position of the probe. The results of these calculations are also used to understand the experimental results obtained by applying the Spernol–Gierer–Wirtz theory to analyze the viscosity dependence of the rotational correlation time of the spin probes. Significant differences are found between 2,2,6,6‐tetramethylpiperidine‐1‐yloxyl and its derivatives containing substituents that are able to form hydrogen bonds with the ionic liquid. Moreover, derivatives substituted with ionic groups at the 4‐position have a large effect on temperature‐induced solvent viscosity, as this is particularly dependent on the nature of the substituent at the 4‐position. These dependencies include the temperature region that can be used to describe interactions between the spin probes and the ionic liquid, diffusion into the free volume during non‐activated (neutral spin probes) and activated (charged spin probes) processes. Additional parameters are the radii of the ionic liquid and the spin probes, which are calculated and measured approximately. In addition, the temperature dependence of the isotropic hyperfine coupling constants of the spin probes results in information about the micropolarity of the ionic liquid. At room temperature, this is comparable to that of the solvent dimethylsulfoxide.  相似文献   

11.
1-Phosphabicyclo[3,3,1]nonane The synthesis of 1-phosphabicyclo[3,3,1]nonane II by free-radical cyclization of the primary phosphine (CH2?CH? CH2)2CH? CH? PH2 I is described. IR data favour a twin-chair conformation of II and its derivatives with oxygen V and sulphur VI . The magnitudes of the 31P NMR chemical shifts and of the phosphorus-hydrogen spin-spin coupling constants (1JPH, 2JPH) are compared with values of aliphatic phosphines R3P. The degree of substitution of CO from Ni(CO)4 by II and other chemical evidences suggest a TOLMAN ligand cone angle of about 120–125°.  相似文献   

12.
Two triangular molybdenum‐tellurium clusters, Mo3Te7‐[S2CN (CH2CH2OH2]3I (1) and Mo3Te4Y3[S2P(iPrO)2]3I (Y3 = 1. 43Te+ 1.57S) (2), were obtained from the reaction of K · HOdtc [HOdtc = S2CN(CH2CH2OH)2] or K·iPr2dtp (iPr2dtp = S2P[OCH(CH3)2]2) with a mixed product of elements Mo, Te(S) and I2 at high temperature. The structures of the two compounds were determined by X‐ray crystallography study. Crystal data are, (1): monoclinic, P21/n, a = 1.6256(3), b = 1.3264(2), c = 1.8808(4) nm, β=96.923°, V = 4.025.9 (14) nm3, Dcal = 3.050g·cm‐3, Z = 4, and (2): monoclinic, P21/n, a = 1.4564(2), b = 2.3917(4), c = 1.5094(3) nm, β = 114.35(2)°, V = 4.0259(14) nm3, Dcal = 2.449 g·cm?3 and Z = 4. Single crystal analyses show that 1 consists of discrete Mo3Te7[S2CN(CH2CH2OH)2]3I connected into three‐dimensional framework via hydrogen bonds, while 2 forms a linear chain via Te(S)—I interaction.  相似文献   

13.
The nonrelativistic and four-component fully relativistic calculations of 1H, 15N, 59Co, 103Rh, and 193Ir shielding constants of pentaammineaquacomplexes of cobalt(III), rhodium(III), and iridium(III) were carried out at the density functional theory (DFT) level of theory. The noticeable deshielding relativistic corrections were observed for nitrogen shielding constants (chemical shifts), whereas those corrections were found to be negligible for protons. For the transition metals cobalt, rhodium, and iridium, relativistic corrections to their nuclear magnetic resonance (NMR) shielding constants were found to be rather small for cobalt and rhodium (some 5–10%), whereas they are essentially larger for iridium (up to 70%).  相似文献   

14.
195Pt NMR chemical shifts of octahedral Pt(IV) complexes with general formula [Pt(NO3)n(OH)6 ? n]2?, [Pt(NO3)n(OH2)6 ? n]4 ? n (n = 1–6), and [Pt(NO3)6 ? n ? m(OH)m(OH2)n]?2 + n ? m formed by dissolution of platinic acid, H2[Pt(OH)6], in aqueous nitric acid solutions are calculated employing density functional theory methods. Particularly, the gauge‐including atomic orbitals (GIAO)‐PBE0/segmented all‐electron relativistically contracted–zeroth‐order regular approximation (SARC–ZORA)(Pt) ∪ 6–31G(d,p)(E)/Polarizable Continuum Model computational protocol performs the best. Excellent second‐order polynomial plots of δcalcd(195Pt) versus δexptl(195Pt) chemical shifts and δcalcd(195Pt) versus the natural atomic charge QPt are obtained. Despite of neglecting relativistic and spin orbit effects the good agreement of the calculated δ 195Pt chemical shifts with experimental values is probably because of the fact that the contribution of relativistic and spin orbit effects to computed σiso 195Pt magnetic shielding of Pt(IV) coordination compounds is effectively cancelled in the computed δ 195Pt chemical shifts, because the relativistic corrections are expected to be similar in the complexes and the proper reference standard used. To probe the counter‐ion effects on the 195Pt NMR chemical shifts of the anionic [Pt(NO3)n(OH)6 ? n]2? and cationic [Pt(NO3)n(OH2)6 ? n]4 ? n (n = 0–3) complexes we calculated the 195Pt NMR chemical shifts of the neutral (PyH)2[Pt(NO3)n(OH)6 ? n] (n = 1–6; PyH = pyridinium cation, C5H5NH+) and [Pt(NO3)n(H2O)6 ? n](NO3)4 ? n (n = 0–3) complexes. Counter‐anion effects are very important for the accurate prediction of the 195Pt NMR chemical shifts of the cationic [Pt(NO3)n(OH2)6 ? n]4 ? n complexes, while counter‐cation effects are less important for the anionic [Pt(NO3)n(OH)6 ? n]2? complexes. The simple computational protocol is easily implemented even by synthetic chemists in platinum coordination chemistry that dispose limited software availability, or locally existing routines and knowhow. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

15.
The reaction of the NHC–disilicon(0) complex [(IAr)Si=Si(IAr)] ( 1 , IAr=:C{N(Ar)C(H)}2, Ar=2,6‐i Pr2C6H3) with two equiv of elemental Te in toluene at room temperature for three days afforded a mixture of the first dimeric NHC–silicon monotelluride [(IAr)Si=Te]2 ( 2 ) and its isomeric complex [(IAr)Si(μ‐Te)Si(IAr)=Te] ( 3 ). When the same reaction was performed for ten days, only 3 was isolated from the reaction mixture. Compound 1 reacted with four equiv of elemental Te in toluene for four weeks, which proceeded through the formation of 2 , 3 and the NHC–disilicon tritelluride complex [{(IAr)Si(=Te)}2Te] ( 5‐Te ), to form the dimeric NHC–silicon ditelluride [(IAr)Si(=Te)(μ‐Te)]2 ( 4 ). The reactions are in line with theoretical mechanistic studies for the formation of 4 . Compound 3 reacted with one equiv of elemental sulfur in toluene to form the first NHC–disilicon sulfur ditelluride complex [{(IAr)Si(=Te)}2S] ( 5‐S ).  相似文献   

16.
A systematic study of the accuracy factors for the computation of 15N NMR chemical shifts in comparison with available experiment in the series of 72 diverse heterocyclic azines substituted with a classical series of substituents (CH3, F, Cl, Br, NH2, OCH3, SCH3, COCH3, CONH2, COOH, and CN) providing marked electronic σ‐ and π‐electronic effects and strongly affecting 15N NMR chemical shifts is performed. The best computational scheme for heterocyclic azines at the DFT level was found to be KT3/pcS‐3//pc‐2 (IEF‐PCM). A vast amount of unknown 15N NMR chemical shifts was predicted using the best computational protocol for substituted heterocyclic azines, especially for trizine, tetrazine, and pentazine where experimental 15N NMR chemical shifts are almost totally unknown throughout the series. It was found that substitution effects in the classical series of substituents providing typical σ‐ and π‐electronic effects followed the expected trends, as derived from the correlations of experimental and calculated 15N NMR chemical shifts with Swain–Lupton's F and R constants.  相似文献   

17.
Synthesis and Properties of Tetrakis(Perfluoroalkyl)Tellurium Te(Rf)4 (Rf = CF3, C2F5, C3F7, C4F9) Te(CF3)4 is obtained from the reaction of Te(CF3)Cl2 with Cd(CF3)2 complexes as a complex with e. g. CH3CN, DMF. It is a light and temperature sensitive hydrolysable liquid. The reaction with fluorides yields the complex anion [Te(CF3)4F], with fluoride ion acceptors the complex cation [Te(CF3)3]+. With traces of water an acidic solution is formed. Te(CF3)4 acts as a trifluoromethylation reagent. The reaction with XeF2 gives hints for the formation of Ye(CF3)4F2. Properties and NMR spectra are discussed. The much more stable complexes of Te(Rf)4 (Rf = C2F5, C3F7, C4F9) are formed from the reaction of TeCl4 with the corresponding Cd(Rf)2 complexes.  相似文献   

18.
Ab initio molecular dynamics (MD) and relativistic density functional NMR methods were applied to calculate the one‐bond Hg? C NMR indirect nuclear spin–spin coupling constants (J) of [Hg(CN)2] and [CH3HgCl] in solution. The MD averages were obtained as J(199Hg? 13C)=3200 and 1575 Hz, respectively. The experimental Hg? C spin–spin coupling constants of [Hg(CN)2] in methanol and [CH3HgCl] in DMSO are 3143 and 1674 Hz, respectively. To deal with solvent effects in the calculations, finite “droplet” models of the two systems were set up. Solvent effects in both systems lead to a strong increase of the Hg? C coupling constant. From a relativistic natural localized molecular orbital (NLMO) analysis, it was found that the degree of delocalization of the Hg 5dσ nonbonding orbital and of the Hg? C bonding orbital between the two coupled atoms, the nature of the trans Hg? C/Cl bonding orbital, and the s character of these orbitals, exhibit trends upon solvation of the complexes that, when combined, lead to the strong increase of J(Hg? C).  相似文献   

19.
In this paper, we have investigated the cumulative peculiarity of the “heavy atom on light atom” effect on the 13C NMR chemical shifts, initiated by the adjacent chalcogens. For this purpose, the most accurate hybrid computational scheme for the calculation of chemical shifts of carbon nuclei, directly bonded with several heavy chalcogens, is introduced and attested on the representative series of molecules. The best hybrid scheme combines the nonrelativistic coupled cluster‐based approach with the different types of corrections, including vibrational, solvent, and relativistic. The dependences of the total relativistic corrections to carbon shielding constants in 2 series of model compounds, namely, X═13C═Y (X, Y = O, S, Se, Te) and C(XH)m(YH)n(ZH)p(QH)sH1‐mH1‐nH1‐pH1‐s (X, Y, Z, Q = S, Se, Te and m, n, p, s = 0, 1), on the total atomic number of the adjacent chalcogens have been obtained.  相似文献   

20.
While exploring the chemistry of tellurium‐containing dichalcogenidoimidodiphosphinate ligands, the first all‐tellurium member of a series of related square‐planar EII(E′)4 complexes (E and E′ are group 16 elements), namely bis(P,P,P′,P′‐tetraphenylditelluridoimidodiphosphinato‐κ2Te,Te′)tellurium(II) (systematic name: 2,2,4,4,8,8,10,10‐octaphenyl‐1λ3,5,6λ4,7λ3,11‐pentatellura‐3,9‐diaza‐2λ5,4λ5,8λ5,10λ5‐tetraphosphaspiro[5.5]undeca‐1,3,7,9‐tetraene), C48H40N2P4Te5, was obtained unexpectedly. The formally TeII centre is situated on a crystallographic inversion centre and is Te,Te′‐chelated to two anionic [(TePPh2)2N] ligands in an anti conformation. The central TeII(Te)4 unit is approximately square planar [Te—Te—Te = 93.51 (3) and 86.49 (3)°], with Te—Te bond lengths of 2.9806 (6) and 2.9978 (9) Å.  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