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 62 毫秒
1.
A series of stable organosuperbases, N‐alkyl‐ and N‐aryl‐1,3‐dialkyl‐4,5‐dimethylimidazol‐2‐ylidene amines, were efficiently synthesized from N,N′‐dialkylthioureas and 3‐hydroxy‐2‐butanone and their basicities were measured in acetonitrile. The derivatives with tert‐alkyl groups on the imino nitrogen were found to be more basic than the tBu P1 (pyrr) phosphazene base in acetonitrile. The origin of the high basicity of these compounds is discussed. In acetonitrile and in the gas phase, the basicity of the alkylimino derivatives depends on the size of the substituent at the imino group, which influences the degree of aromatization of the imidazole ring, as measured by 13C NMR chemical shifts or by the calculated ΔNICS(1) aromaticity parameters, as well as on solvation effects. If a wider range of imino‐substituents, including electron‐acceptor substituents, is treated in the analysis then the influence of aromatization is less predominant and the gas‐phase basicity becomes more dependent on the field‐inductive effect, polarizability, and resonance effects of the substituent.  相似文献   

2.
The potential applications of tetrel bonds involving π‐molecules in crystal materials and biological systems have prompted a theoretical investigation of the strength of π···σ‐hole tetrel bond in the systems with acetylene and its derivatives of CH3, AuPH3, Li, and Na as well as benzene as the π electron donors. A weak tetrel bond (ΔE < 15 kJ/mol) is found between acetylene and tetrel donor molecule TH3F (T = C, Si, Ge, Sn, and Pb). All substituents strengthen the π tetrel bond, but the electron‐donating sodium atoms have the largest enhancing effect and the interaction energy is up to about 24 kJ/mol in C2Na2‐CH3F. The electron‐donating ability of the AuPH3 fragment is intermediate between the methyl group and alkali metal atom. The origin of the stability of the π tetrel‐bonded complex is dependent on the nature of the tetrel donor and acceptor molecules and can be regulated by the substituents.  相似文献   

3.
The 1H NMR spectra of 35 cyclic and acyclic esters are analysed to give the 1H chemical shifts and couplings. The substituent chemical shifts of the ester group were analysed using three‐bond (γ) effects for near protons and the electric field, magnetic anisotropy and steric effect of the ester group for more distant protons. The electric field is calculated from the partial atomic charges on the O?C = O atoms, and the asymmetric magnetic anisotropy of the carbonyl group acts at the midpoint of the C = O bond. The values of the anisotropies Δχparl and Δχperp were for the aliphatic esters 10.35 and ?18.84 and for the conjugated esters 7.33 and ?15.75 (×10?6 Å3/molecule). The oxygen steric coefficients found were 104.4 (aliphatic C = O), 45.5 (aromatic C = O) and 16.0 (C–O) (×10?6 Å6/molecule). After parameterisation, the overall RMS error for the data set of 280 entries was 0.079 ppm. The strongly coupled 1H NMR spectra of the 2‐methyl, 3‐methyl and 4‐methyl γ‐butyrolactones were analysed and the methyl conformational equilibrium obtained from the observed couplings. The observed versus calculated density functional theory (DFT) ΔG(ax‐eq) was 1.0 (1.01), 0.34 (0.54) and 0.65 (0.71) kcal/mol res. The shielding effect of a methyl cis to a proton in the five‐membered lactone rings is ?0.40 ±0.05 ppm and deshielding trans effect 0.12 ±0.05 ppm, which is common to both five and six membered rings. The cis/trans isomerism in the vinyl esters methyl acrylate, crotonate and methacrylate and methyl furoate was examined using the 1H chemical shifts. The calculated shifts of both the cis and trans isomers were in good agreement with the observed shifts. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

4.
We present J=0 calculations of all bound and pseudobound vibrational states of Li3 in its first‐excited electronic doublet state by using a realistic double many‐body expansion potential‐energy surface and a minimum‐residual filter diagonalization technique. The action of the system Hamiltonian on the wave function was evaluated by the spectral transform method in hyperspherical coordinates. Calculations of the vibrational spectra were carried out both without consideration and with consideration of geometric‐phase effects. Dynamic Jahn–Teller and geometric‐phase effects are found to play a significant role, while the calculated fundamental symmetric stretching frequency is larger by 8.3% than its reported experimental value of 326 cm−1. From the neighbor‐spacing distributions of the levels, it is observed that the title vibrational spectrum is quasiregular in the short range and quasi‐irregular in the long range. By the Δ2 standard defined in this article, it is found that the spectra are more nonuniform than those of the “trough” states for the ground electronic state. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 75: 89–109, 1999  相似文献   

5.
Mass‐analyzed threshold ionization spectra of jet‐cooled [(η6‐PhMe)(η6‐PhH)Cr] and [(η6‐Ph2)(η6‐PhH)Cr] reveal with unprecedented accuracy the effects of methyl and phenyl groups on the electronic structure of bis(η6‐benzene)chromium. These “pure” substituent effects allow quantitative experimental determination of the ionization energy changes caused by the mutual substituent influence in bisarene systems. Two types of such influence have been revealed for the first time in bis(η6‐toluene)chromium.  相似文献   

6.
The capability of a gaseous Brønsted acid HB to deliver protons to a base is usually described by the gas‐phase acidity (GA) value of the acid. However, GA values are standard Gibbs energy differences and refer to individual gas pressures of 1 bar for acid HB, base B?, and proton H+. We show that the GA value is not suited to describe the bulk acidity of a gaseous acid. Here the pressure dependence of the activities of HB, H(HB)n+, and B(HB)m? that result from gaseous autoprotolysis have to be considered. In this work, the pressure‐dependent absolute chemical potential of the proton in the representative gaseous proton acids CH4, NH3, H2O, HF, and HCl was worked out and the general theory to describe bulk gas phase acidity—that can directly be compared with solution acidity—was developed.  相似文献   

7.
Various recent computational studies initiated this systematic re‐investigation of substituent effects on aromatic edge‐to‐face interactions. Five series of Tröger base derived molecular torsion balances (MTBs), initially introduced by Wilcox and co‐workers, showing an aromatic edge‐to‐face interaction in the folded, but not in the unfolded form, were synthesized. A fluorine atom or a trifluoromethyl group was introduced onto the edge ring in ortho‐, meta‐, and para‐positions to the C?H group interacting with the face component. The substituents on the face component were varied from electron‐donating to electron‐withdrawing. Extensive X‐ray crystallographic data allowed for a discussion on the conformational behavior of the torsional balances in the solid state. While most systems adopt the folded conformation, some were found to form supramolecular intercalative dimers, lacking the intramolecular edge‐to‐face interaction, which is compensated by the gain of aromatic π‐stacking interactions between four aryl rings of the two molecular components. This dimerization does not take place in solution. The folding free enthalpy ΔGfold of all torsion balances was determined by 1H NMR measurements by using 10 mM solutions of samples in CDCl3 and C6D6. Only the ΔGfold values of balances bearing an edge‐ring substituent in ortho‐position to the interacting C?H show a steep linear correlation with the Hammett parameter (σmeta) of the face‐component substituent. Thermodynamic analysis using van′t Hoff plots revealed that the interaction is enthalpy‐driven. The ΔGfold values of the balances, in addition to partial charge calculations, suggest that increasing the polarization of the interacting C?H group makes a favorable contribution to the edge‐to‐face interaction. The largest contribution, however, seems to originate from local direct interactions between the substituent in ortho‐position to the edge‐ring C?H and the substituted face ring.  相似文献   

8.
刘志宏  赵莉  胡满成 《中国化学》2003,21(12):1569-1572
Introduction   2MgO·B2 O3(Mg2 B2 O5)and 2MgO·B2 O3·H2 Omightbepreparedaswhiskermaterials .12MgO·B2 O3·H2 OnamedszaibelyiteisamagnesiumboratemineralwithastructuralformulaofMg2 [B2 O4 (OH) 2 ].2 Itisdifficulttosynthesizethiscompoundinthelaboratory .Recently ,weobtainedasimilarcompound 2MgO·B2 O3·1 5H2 Owhenwetriedtopreparewhiskerof 2MgO·B2 O3·H2 Obythephasetransformationof 2MgO·2B2 O3·MgCl2 ·14H2 OinH3BO3solutionunderhydrothermalcondition .Itishope fultopreparewh…  相似文献   

9.
The gas‐phase bond‐dissociation energies of a SO2–imidazolylidene leaving group of three gold(I) benzyl imidazolium sulfone complexes are reported (E0=46.6±1.7, 49.6±1.7, and 48.9±2.1 kcal mol?1). Although these energies are similar to each other, they are reproducibly distinguishable. The energy‐resolved collision‐induced dissociation experiments of the three [L]–gold(I) (L=ligand) carbene precursor complexes were performed by using a modified tandem mass spectrometer. The measurements quantitatively describe the structural and electronic effects a p‐methoxy substituent on the benzyl fragment, and trans [NHC] and [P] gold ligands, have towards gold carbene formation. Evidence for the formation of the electrophilic gold carbene in solution was obtained through the stoichiometric and catalytic cyclopropanation of olefins under thermal conditions. The observed cyclopropane yields are dependent on the rate of gold carbene formation, which in turn is influenced by the ligand and substituent. The donation of electron density to the carbene carbon by the p‐methoxy benzyl substituent and [NHC] ligand stabilizes the gold carbene intermediate and lowers the dissociation barrier. Through the careful comparison of gas‐phase and solution chemistry, the results suggest that even gas‐phase leaving‐group bond‐dissociation energy differences of 2–3 kcal mol?1 enormously affect the rate of gold carbene formation in solution, especially when there are competing reactions. The thermal decay of the gold carbene precursor complex was observed to follow first‐order kinetics, whereas cyclopropanation was found to follow pseudo‐first‐order kinetics. Density‐functional‐theory calculations at the M06‐L and BP86‐D3 levels of theory were used to confirm the observed gas‐phase reactivity and model the measured bond‐dissociation energies.  相似文献   

10.
Motivated by the need for chemical strategies designed to tune peptide fragmentation to selective cleavage reactions, benzyl ring substituent influence on the relative formation of carbocation elimination (CCE) products from peptides with benzylamine‐derivatized lysyl residues has been examined using collision‐induced dissociation (CID) tandem mass spectrometry. Unsubstituted benzylamine‐derivatized peptides yield a mixture of products derived from amide backbone cleavage and CCE. The latter involves side‐chain cleavage of the derivatized lysyl residue to form a benzylic carbocation [C7H7]+ and an intact peptide product ion [(MHn)n+ – (C7H7)+](n‐1)+. The CCE pathway is contingent upon protonation of the secondary ε‐amino group (Nε) of the derivatized lysyl residue. Using the Hammett methodology to evaluate the electronic contributions of benzyl ring substituents on chemical reactivity, a direct correlation was observed between changes in the CCE product ion intensity ratios (relative to backbone fragmentation) and the Hammett substituent constants, σ, of the corresponding substituents. There was no correlation between the substituent‐influenced gas‐phase proton affinity of Nε and the relative ratios of CCE product ions. However, a strong correlation was observed between the π orbital interaction energies (ΔEint) of the eliminated benzylic carbocation and the logarithm of the relative ratios, indicating the predominant factor in the CCE pathway is the substituent effect on the level of hyperconjugation and resonance stability of the eliminated benzylic carbocation. This work effectively demonstrates the applicability of σ (and ΔEint) as substituent selection parameters for the design of benzyl‐based peptide‐reactive reagents which tune CCE product formation as desired for specific applications. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
Quantum chemical calculations have been performed to study the single‐electron halogen bonds in HO···ClF and HS···ClF complexes. The calculation methods have a larger effect on the S···Cl halogen bond than on the O···Cl one. The interaction strength in HO···ClF complex is stronger than that in HS···ClF one, but the presence of methyl group in the halogen acceptor makes the sequence reverse. The methyl group has a greater effect on the S···Cl halogen bond than on the O···Cl one. The charge analyses indicate that the methyl group is electron‐donating and the electron‐donating role in the H3CS? ClF complex is larger than in the H3CO? ClF one. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

12.
A novel 19F NMR‐based method for monitoring the enzymatic oxidation of thia fatty acid analogues is presented. Our approach is based on the observation that methyl ω‐monofluorinated 9‐thia‐ and 10‐thiaoctadecanoates and their S‐oxide and S‐dioxide derivatives are easily distinguishable via their 1H‐decoupled 19F spectra. These long‐range substituent effects were used to probe the regio‐ and chemoselectivity of stearoyl ACP (acyl carrier protein) Δ9 desaturase‐mediated sulfoxidation. The results clearly demonstrate that mono‐oxygenation of a 10‐thia analogue ACP Δ9 desaturase was more efficient than that of a 9‐thia substrate. A product previously undetected by TLC was observed for the first time in the product mixture obtained from 18‐fluoro‐9‐thiaoctadecanoyl‐ACP. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

13.
A theoretical study of the C?H···N hydrogen bond in the interactions of trihalomethanes CHX3 (X = F, Cl, Br) with ammonia and its halogen derivatives NH2Y (Y = F, Cl, Br) has been carried out thoroughly. The complexes are quite stable, and their stability increases in going from CHF3 to CHCl3 then to CHBr3 when Y keeps unchanged. With the same CHX3 proton donor, enhancement of the gas phase basicity of NH2Y strengthens stability of the CHX3···NH2Y complex. The C?H···N hydrogen bond strength is directly proportional to the increase of proton affinity (PA) at N site of NH2Y and the decrease of deprotonation enthalpy (DPE) of C?H bond in CHX3. The CHF3 primarily appears to favor blue shift while the red‐shift is referred to the CHBr3. The blue‐ or red‐shift of CHCl3 strongly depends on PA at N site of NH2Y. We suggest the ratio of DPE/PA as a factor to predict which type of hydrogen bond is observed upon complexation. The SAPT2+ results show that all C?H···N interactions in the complexes are electrostatically driven regardless of the type of hydrogen bond, between 48% and 61% of the total attractive energy, and partly contributed by both induction and dispersion energies.  相似文献   

14.
On the basis of a comparison of chemical shifts and wavenumbers of several secondary thioamides and amides having monocationic substituents attached to thiocarbamoyl or carbamoyl groups by a polymethylene chain, new intramolecular unconventional N···H+···N hydrogen bonding effects were discovered. It is argued that the CH2—N rotation is hindered and two +H···NHCH3 non‐equivalent protons occur in a proton spectrum of hydrochloride 1a (at 10.68 and 2.77 ppm, respectively) instead of two +NH2CH3 protons. Presumably, the above steric factors inhibit the acidic hydrolysis of 1a (stabilized by strong intramolecular N···H+···N hydrogen bonds) to an amide and prevent intramolecular cyclization of 2a (stabilized by strong intramolecular neutral–neutral N···HN hydrogen bonds) to a cyclic amidine. Postulation of additional dihydrogen bond formation is helpful in understanding the spectroscopic differences of 4 and 5 . The above new bonding is also compared with intramolecular N···H—N+ hydrogen bonds in primary amine salts 7 and 8 . In contrast to 3 , a cooperative hydrogen bonded system is observed in 9 and 10 . The weak hydrogen bonds in 7 – 10 facilitate the hydrolysis and cyclization reactions of secondary thioamides. The spectroscopic data for secondary (thio)amides are especially useful for characterizing the electronic situation at the (thio)carbamoyl nitrogen atoms and they are perfectly correlated with the reactivity. Examples of chelation of protons by thioamides ( 11 and 12 ), which contain strongly electron‐donating pyrimidine groups, are presented to show the contribution of dihydrogen bonding in the protonation reaction similar to 1 and 4 . Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

15.
The reaction of Cs3As7 with diphenylacetylene in the presence of 18 crown‐6 in liquid ammonia results in the formation of the new compound [Cs( 18 crown‐6)]2As7C14H11 · 6NH3, which crystallizes in black monoclinic crystals. It contains the first monosubstituated heptaarsenide anion with a hydrocarbon‐only substituent and theoretical calculations show a significant influence of the organic substituent on the electronic structure within the cage. The (Z)‐1, 2‐diphenylethenyl‐heptaarsenide di‐anion can be seen as the first step towards the formation of 1, 2,3‐triarsolides. Further experiments regarding the reaction of Rb3As11 and Cs3As11 with acetylene gas in liquid ammonia reveal the formation of the diarsabarrelene As2C6H6, which crystallizes as colorless orthorhombic crystals. Calculations based on the structural data obtained by X‐ray crystallography show the electronically inert character of the arsenic lone pair.  相似文献   

16.
Two series of 4‐substituted N‐[1‐(pyridine‐3‐ and ‐4‐yl)ethylidene]anilines have been synthesized using different methods of conventional and microwave‐assisted synthesis, and linear free‐energy relationships have been applied to the 13C NMR chemical shifts of the carbon atoms of interest. The substituent‐induced chemical shifts have been analyzed using single substituent parameter and dual substituent parameter methods. The presented correlations describe satisfactorily the field and resonance substituent effects having similar contributions for C1 and the azomethine carbon, with exception of the carbon atom in para position to the substituent X. In both series, negative ρ values have been found for C1′ atom (reverse substituent effect). Quantum chemical calculations of the optimized geometries at MP2/6‐31G++(d,p) level, together with 13C NMR chemical shifts, give a better insight into the influence of the molecular conformation on the transmission of electronic substituent effects. The comparison of correlation results for different series of imines with phenyl, 4‐nitrophenyl, 2‐pyridyl, 3‐pyridyl, 4‐pyridyl group attached at the azomethine carbon with the results for 4‐substituted N‐[1‐(pyridine‐3‐ and ‐4‐yl)ethylidene]anilines for the same substituent set (X) indicates that a combination of the influences of electronic effects of the substituent X and the π1‐unit can be described as a sensitive balance of different resonance structures.  相似文献   

17.
A2B‐type B‐methoxy subporphyrins 3 a – g and B‐phenyl subporphyrins 7 a – c , e , g bearing meso‐(2‐substituted)aryl substituents are synthesized, and their rotational dynamics are examined through variable‐temperature (VT) 1H NMR spectroscopy. In these subporphyrins, the rotation of meso‐aryl substituents is hindered by a rationally installed 2‐substituent. The rotational barriers determined are considerably smaller than those reported previously for porphyrins. Comparison of the rotation activation parameters reveals a variable contribution of ΔH and ΔS in ΔG. 2‐Methyl and 2‐ethyl groups of the meso‐aryl substituents in subporphyrins 3 e , 3 f , and 7 e induce larger rotational barriers than 2‐alkoxyl substituents. The rotational barriers of 3 g and 7 g are reduced by the presence of the 4‐dibenzylamino group owing to its ability to stabilize the coplanar rotation transition state electronically. The smaller rotational barriers found for B‐phenyl subporphyrins than for B‐methoxy subporphyrins indicate a negligible contribution of SN1‐type heterolysis in the rotation of meso‐aryl substituents.  相似文献   

18.
The title compound [alternative name: 8‐methyl‐8‐aza­bi­cyclo­[3.2.1]­octan‐3‐yl­methyl p‐toluene­sulfonate(N8B)–bor­ane], C16H26BNO3S, has the tosyl­oxy­methyl substituent in an endo position. The BH3 group is equatorial and the (N‐bonded) methyl group is axial, relative to the six‐membered heterocycle. The N—B bond of 1.649 (8) Å is one of the longest known.  相似文献   

19.
The non‐ionic europium(III) complex [Eu(TTA)3·DAF]·0.5C7H8 (TTA = 2‐thenoytrifluoroacetonate, DAF = 4, 5‐diazafluoren‐9‐one) was synthesized. The structural determination has been carried out. DAF coordination induces the both excitation spectra in the solid state and solution having a red shift and sensitizes Eu3+ luminescence under visible light excitation.  相似文献   

20.
Single crystals of AlBr3 · NH3 and AlI3 · NH3 sufficient in size for X‐ray structure determinations were obtained by evaporation/ sublimation of the respective compound from its melt. The ammoniates were synthesized by the reaction of the pure halide with NH3 at ‐78°C and following homogenization by slowly heating the reaction mixture up to the melting points of the ammoniates (124°C and 126°C, respectively). The X‐ray structure determinations for both monoammoniates were successfully carried out for the heavy atom positions (no hydrogen atoms): AlBr3 · NH3: Pbca, Z = 16, a = 11.529 (5) Å, b = 12.188 (2) Å, c = 19.701 (4) Å AlI3 · NH3: Pbca, Z = 8, a = 13.536 (5) Å, b = 8.759 (2) Å, c = 14.348 (4) Å The structures contain tetrahedral molecules Al(NH3)X3 with X = Br, I. They are not isotypic. The main difference is given for the coordination of NH3 by X from neighbouring molecules. In Al(NH3)Br3 one of the two crystallographically independent NH3 ligands has 6Br and the other 7Br as neighbours whereas in Al(NH)3I3 only 5I surround the one kind of NH3.  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