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.
《Polyhedron》1986,5(3):899-900
By using newer structural and NMR data for Mo2(O2CR)4 compounds with R = PhCH(OH) and CH3 the diamagnetic anisotropy of the Mo-Mo quadruple bonds has been estimated as −(210±30) × 10−36 m3 per molecule.  相似文献   

2.
《Polyhedron》1988,7(14):1279-1284
The reactions of Mo2(O2CCH3)2Cl2(PPh3)2 with the bidentate phosphine ligands Ph2PCH2CH2PPh2(dppe), Ph2PCHCHPPh2(dppee) and o-C6H4(PPh2)2(dppbe) yield purple complexes of the stoichiometry Mo2(O2CCH3)2X2(LL) in which the acetate ligands are believed to be cis and the LL ligands bridge the dimolybdenum core so that their P atoms are anti to one another. The relationship of these complexes to the previously reported complex Mo2(O2CCH3)2Cl2(dppm), where dppm = Ph2PCH2PPh2, is discussed. These complexes have been characterized on the basis of their electrochemical redox properties, and electronic absorption and NMR (1H and 31P{1H}) spectral measurements.  相似文献   

3.
 The topological analysis of the electron localization function has been applied to complexes representative of the weak, medium and strong hydrogen bond. For both the weak and the medium hydrogen bonds, the number of basins in the complexes is the sum of those of the moieties. In this case, the formation of a weak or a medium hydrogen-bonded complex does not involve a chemical reaction. In the weak hydrogen bond case, the reduction of the localization domain yields two domains in the first step, which can be partitioned afterwards into valence and core domains. In contrast, for medium complexes the core–valence separation is the first event which occurs during the reduction process and therefore the complex should be considered as a single molecular species. Moreover, the analysis of the basin population variance indicates in this case a noticeable delocalization between the V(A, H) and V(B) basins. Finally, the symmetrical strong hydrogen bond has a protonated basin V(H) at the bond midpoint. Such a topology corresponds to an incomplete proton transfer and to a rather covalent bond. Received: 19 April 1999 / Accepted: 22 July 1999 / Published online: 17 January 2000  相似文献   

4.
《Solid State Sciences》2000,2(3):385-389
The most commonly used equations correlating bond valence and bond length have been critically compared. It has been shown that the Zachariasen equation is more accurate than the Brown–Shannon equation. Doubts already voiced about the universality of the constant B in the Brown–Altermatt equation with a value of 0.37 Å have been hereby confirmed. Moreover, by a method based on the comparison of formal oxidation states and valences of molybdenum in suitable oxides, the parameters relative to the Zachariasen equation have been accurately determined for the MoO bond in MoO6 octahedra. Their values are R1=1.8790 and B=0.3048 Å in the 3–6 v.u. range.  相似文献   

5.
Six dioxetanone molecules, ranging in complexity from simple dioxetanone to firefly dioxetanone, were studied by performing M06/6-311G(d,p) calculations. The quantum theory of atoms in molecules and the electron localization function was applied to analyze the peroxide and carbon–carbon bonds of the dioxetanone ring. Both approaches demonstrated that the peroxide bond is not covalent, but charge-shifted. This means that for this bond the covalent “electron sharing” is relatively unimportant, and it is the stabilizing resonance energy that causes the bonding. For the contrary, the carbon–carbon bond is covalent. These discoveries indicate that no biradical species should be formed in the dioxetanone decomposition, and that the most probable rate-determining step should be the carbon–carbon cleavage.  相似文献   

6.
X-ray diffraction studies reveal that pyrazine-2-thiol undergoes condensation to 2,2′-dithiobispyrazine [systematic name: 2-(pyrazin-2-yldisulfanyl)pyrazine], C8H6N4S2 ( I ), under aerial conditions. In the molecule of I , the pyrazine rings are arranged in an almost perpendicular manner, with an absolute value of the C—S—S—C torsion angle of −91.45 (6)°. A search in the Cambridge Structural Database confirmed that such a conformation is typical for disulfide compounds. Three different rotamers of disulfide I were studied using quantum theoretical studies. The rotamer of lowest energy was observed in the crystalline state in the structure stabilized by hydrogen-bond, chalcogen-bond and stacking interactions. Further quantum chemical computations confirm that 2,2′-dithiobispyrazine can react according to the SN2 mechanism.  相似文献   

7.
An extension of the harmonic oscillator model of aromaticity (HOMA) model for systems with boron–boron bonds is presented. For the first time, the parameters of the HOMA model are estimated using only theoretically calculated bond lengths. The HOMA parameters obtained make geometric aromaticity studies possible for a large number of compounds containing the boron–boron bond. The derived HOMA parameters have been used to investigate how the introduction of the boron–boron moiety in the structure of selected hydrocarbons modifies their aromaticity. The conclusion is that the insertion of a boron–boron bond usually strongly decreases the aromaticity of the boron-containing compounds in comparison to their parent hydrocarbons.  相似文献   

8.
Quantum chemical calculations are used to estimate the equilibrium C? NO bond dissociation energies (BDEs) for eight X? NO molecule (X = CCl3, C6F5, CH3, CH3CH2, iC3H7, tC4H9, CH2CHCH2, and C6H5CH2). These compounds are studied by employing the hybrid density functional theory (B3LYP, B3PW91, B3P86) methods together with 6‐31G** and 6‐311G** basis sets and the complete basis set (CBS‐QB3) method. The obtained results are compared with the available experimental results. It is demonstrated that B3P86/6‐31G** and CBS‐QB3 methods are accurate for computing the reliable BDEs for the X? NO molecule. Considering the inevitably computational cost of CBS‐QB3 method and the reliability of the B3P86 calculations, B3P86 method with 6‐31G** basis set may be more suitable to calculate the BDEs of the C? NO bond. The solvent effects on the BDEs of the C? NO bond are analyzed and it is shown that the C? NO BDEs in a vacuum computed by using B3PW91/6‐311G** method are the closest to the computed values in acetontrile and the average solvent effect is 1.48 kcal/mol. Subsequently, the substituent effects of the BDEs of the C? NO bond are further analyzed and it is found that electron denoting group stabilizes the radical and as a result BDE decreases; whereas electron withdrawing group stabilizes the group state of the molecule and thus increases the BDE from the parent molecule. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2009  相似文献   

9.
We describe the synthesis and preliminary study of two molecules, in which a fluorine atom is positioned proximately above the π-orbitals of a C═C bond or else wherein a C-F bond interacts in a "head on" fashion with a proximate C-H bond. The spectroscopic characteristics of these unusual interactions are documented, X-ray crystallographic analyses are reported, and theoretical calculations are employed to support the observed spectroscopy.  相似文献   

10.
Quantum-chemical calculations of the H2O2 and F2 molecules using different computational schemes, basis sets, and procedures for the inclusion of electron correlation were performed. High-resolution X-ray diffraction study of the electron density distribution in the crystals of 2,5-dimethyl-2,5-dihydroperoxyhexane and 2,5-dimethyl-2,5-dihydroperoxyhex-3-yne was carried out. Joint analysis of the results obtained showed that the formally covalent O—O and F—F bonds correspond to a specific type of interatomic interaction. This type is intermediate between the shared and closed-shell interactions (the latter are typical of the ionic systems and van der Waals molecules).  相似文献   

11.
The interaction between AuOH and the lone-pair donors (HF, H2O) is shown to result in well-bound complexes whose structure resembles that of the corresponding H-bonded systems with the gold atom replacing hydrogen. The dissociation energies are estimated to be 10.7 and 27.4 kcal/mol for HFAu–OH and H2OAu–OH, respectively. However, the interaction between AuOH and the lone pair donors is found to involve significant charge transfer. Furthermore, the Au–O stretching frequency increases upon the complex formation. It is concluded that, in spite of certain similarity to the H-bonded species, the Au-bonded complexes should be considered as Lewis acid–base pairs.  相似文献   

12.
Fast and reliable prediction of bond orders in organic systems based upon experimentally measured quantities can be performed using electron density features at bond critical points (J Am Chem Soc 105:5061–5068, 1983; J Phys Org Chem 16:133–141, 2003; Acta Cryst B 61:418–428, 2005; Acta Cryst B 63:142–150, 2007). These features are outcomes of low-temperature high-resolution X-ray diffraction experiments. However, a time-consuming procedure of gaining these quantities makes the prediction limited. In the present work we have employed an empirical approach AlteQ (J Comput Aided Mol Des 22:489–505, 2008) for evaluation of electron density properties. This approach uses a simple exponential function derived from comparison of electron density, gained from high-resolution X-ray crystallography, and distance to atomic nucleus what allows calculating density distribution in time-saving manner and gives results which are very close to experimental ones. As input data AlteQ accepts atomic coordinates of isolated molecules or molecular ensembles (for instance, protein–protein complexes or complexes of small molecules with proteins, etc.). Using AlteQ characteristics we have developed regression models predicting Cioslowski–Mixon bond order (CMBO) indexes (J Am Chem Soc 113(42):4142–4145, 1991). The models are characterized by high correlation coefficients lying in the range from 0.844 to 0.988 dependently on the type of covalent bond, thereby providing a bonding quantification that is in reasonable agreement with that obtained by orbital theory. Comparative analysis of CMBOs approximated using topological properties of AlteQ and experimental electron densities has shown that the models can be used for fast determination of bond orders directly from X-ray crystallography data and confirmed that AlteQ characteristics can replace experimental ones with satisfactory extent of accuracy.  相似文献   

13.
The energetics of the phenol O–H bond in methanol and the water O–H bond in liquid water were investigated by microsolvation modelling and statistical mechanics Monte Carlo simulations. The microsolvation approach was based on density functional theory calculations. Optimised structures for clusters of phenol and the phenoxy radical with one and two methanol molecules are reported. By analysing the differential solvation of phenol and the phenoxy radical in methanol, we predict that the phenol O–H homolytic bond dissociation enthalpy in solution is 24.3±11 kJ/mol above the gas-phase value. The analysis of the water O–H bond dissociation by microsolvation was based on optimised structures of OH–(H2O)1–6 and –(H2O)1–7 clusters. Microsolvation modelling and statistical mechanics simulations predict that the HO–H bond dissociation enthalpies in the gas phase and in liquid water are very similar. Our results stress the importance of estimating the differences between the solvation enthalpies of the radical species and the parent molecule and the limitations of local models based on microsolvation.Proceedings of the 11th International Congress of Quantum Chemistry satellite meeting in honor of Jean-Louis Rivail  相似文献   

14.
Mo Y 《Organic letters》2006,8(3):535-538
[structure: see text]. The block-localized wave function (BLW) method has been employed to probe the origin of the very short linking C-C bond (1.436 A) in tetrahedranyltetrahedrane. Computations show that the vicinal hyperconjugative interactions between the two tetrahedranyl groups is stronger than the conjugation in butadiene, and if there were no hyperconjugation effect, the bond distance would be 1.491 A. Thus, both the hybridization mode and hyperconjugative interactions contribute to the shortening of the central C-C bond in tetrahedranyltetrahedrane.  相似文献   

15.
Of the several hundred examples of transition metal dihydrogen complexes that have been reported to date, the vast majority have H-H distances of less than 1.0 Angstrom. A small number of complexes have been reported with distances in the range of 1.1 to 1.5 Angstrom. These complexes have been termed elongated dihydrogen complexes. In this review, experimental methods for structure determination of such complexes are summarized, along with computational approaches which have proven useful in understanding the structures of these molecules.  相似文献   

16.
《Tetrahedron letters》1988,29(21):2551-2554
Participation of the CSn bond in the development of beta positive charge in aqueous trifluoroethanol does not depend on solvent nucleophilicity but does depend on SnCCX stereochemistry. These observations favor a hyperconjugative interaction and predict a phenomenal acceleration in the antiperiplanar arrangement.  相似文献   

17.
《Tetrahedron》1988,44(17):5495-5506
In this report, we describe three possibilities for aminoacid synthesis using an enantioselective electrophilic process. Thus, enantioselective carboxylation, alkylation and protonation of Schiff bases yield optically active aminoacids with e.e. up to 76%.  相似文献   

18.
The variation of the An–Cl bond distance in ground-state actinide trichloride (AnCl3) molecules has been studied by density functional theory calculations using the B3LYP exchange–correlation functional in conjunction with small-core relativistic energy-consistent pseudopotentials for the actinides. The ground electronic states and the ground-state molecular properties of the trichlorides of heavy actinides (An = Bk–Lr) are reported in this paper the first time. Extending the present results with literature data on the light actinide trichlorides (AnCl3, An = Th–Cm), the trend in the bond distance has been evaluated for the whole actinide row. The contraction is well manifested in the major part of the actinide row (An = U–Fm). The deviations at the beginning (Th, Pa) and end of the row (Md, No) have been explained by minor differences in the bonding interactions.  相似文献   

19.
To investigate the π bonding electron contribution to N-HO[double bond, length as m-dash]C hydrogen-bond (H-bond) formation, we applied IR cavity ringdown spectroscopy to jet-cooled pyrrole-acetone (Py-Ac) binary clusters. The observed NH stretching vibrations were analyzed by density functional theory (DFT), in which the energetically optimized structures, harmonic frequencies, and interaction energies were calculated for various sizes of binary clusters. We observed three NH stretching vibrations, ascribed to binary clusters at 3406, 3388, and 3335 cm(-1). These were assigned to H-bonded NH stretches of the Py(2)-Ac(1), Py(1)-Ac(1), and Py(1)-Ac(2) clusters, respectively. The Py(1)-Ac(1) cluster has a single N-HO[double bond, length as m-dash]C H-bonded structure with C(s) symmetry, while the Py(1)-Ac(2) cluster has a cyclic structure formed by a single N-HO[double bond, length as m-dash]C H-bond, dipole-dipole interactions, and weak CH H-bonds. A natural bond orbital (NBO) analysis was performed to reveal the H-bond strength in Py-Ac binary clusters. For the Py(1)-Ac(2) cluster, we found that the donor-acceptor interactions are not only the n →σ* type (O atom lone pair to the NH anti-bonding orbitals), but also the π→σ* type (the CO π bonding to the NH anti-bonding orbitals). By analyzing the relationship between the frequency shift and the stabilization energy in donor-acceptor interactions, we concluded that larger red-shift of the NH stretching vibration in the Py(1)-Ac(2) can be explained by not only the lone pair and the π electron contributions to the N-HO[double bond, length as m-dash]C H-bond, but also the dipole-interaction between Py and non-H-bonded Ac. We also discussed the structures of Py(2)-Ac(1) clusters.  相似文献   

20.
The specific cleavage of N-Cα bonds on the peptide backbone to form the so-called ‘c’ and ‘z + 2’ products, which can be used for the rapid determination of protein amino-acid sequences, has been examined to clarify the mechanism(s) that occur during hydrogen abstraction induced by bombardment with 337-nm laser photons in matrix-assisted laser desorption/ionization (MALDI) method. Intramolecular hydrogen abstraction, which results from the hydrogen(s) on the Cα or Cβ carbon, did not occur with a deuterium-labeled dodecapeptide. To confirm a proposition that intermolecular hydrogen abstraction occurs between the peptide and the MALDI matrix, a deuterium dodecapeptide embedded in a deuterium 2,5-dihydroxybenzoic acid matrix at a molar ratio of 1:7000 was analyzed. The resulting deuterium c product ions suggested that c ions form via intermolecular hydrogen abstraction, although the results obtained did not deny any other possibilities such as intramolecular transfer of labile hydrogen. A mechanism for the N-Cα bond cleavage has been proposed that the formation of hypervalent radical species and subsequent prompt bond cleavages occur. The proposed mechanism successfully rationalizes the formation of both the z + 2 and the c product ions.  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