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 46 毫秒
1.
As a continuation of our systematic investigation of the effect of substituents on the ring geometry and dynamics in silacyclobutanes and in order to explore the role of the silicon atom as a mediator for electronic interactions between the attached fragments, we studied the molecular structure of 1,1-diethynylsilacyclobutane (DESCB) by means of gas-phase electron diffraction and ab initio calculations. The structural refinement of the electron diffraction data yielded the following bond lengths (ra) and bond angles (uncertainties are 3σ): r(Si–C)=1.874(2) Å, r(Si–C)=1.817(1) Å, (C–Si–C)=79.2(6)°, (C–Si–C)=106.5(6)°. The geminal Si–CC moieties were found to be bent outwards by 3.1(15)° and the puckering angle was determined to be 30.0(15)°. The evidently short Si–C bond length, which was also reproduced by the ab initio calculations, could be rationalized as being the consequence of the electronic interaction between the outer π charges of the triple bond and the 3pπ orbitals at the silicon atom. It is also likely that the conjugation of the geminal ethynyl groups leads to an enhancement of this bond contraction. Electrostatic interactions and the subsequent reduction of the covalent radius of the silicon atom may also contribute to this bond shortening. It has been found that the endocyclic Si–C bond length fits nicely within a scheme describing a monotonous decrease of the Si–C bond length with the increase of the electronegativity of the substituent in various geminally substituted silacyclobutanes.A series of related silacyclobutanes and acyclic diethynylsilanes have been studied by applying various ab initio methods and their optimized structures were compared to the structure of DESCB. Among these compounds are 1,1-dicyanosilacyclobutane (DCYSCB), which is isoelectronic to DESCB, 1,1-diethynylcyclobutane (DECB) which is isovalent to DESCB, monoethynylsilacyclobutane (MESCB) and monocyanosilacyclobutane (MCYSCB). Searching for reasonable support for the explanation of the structural results of DESCB we performed detailed natural population analysis as well as Mulliken population analysis (MPA) on DESCB and other related molecules. In contrast to the Mulliken charges, the natural atomic charges provided helpful information concerning the bonding properties in DESCB and the corresponding compounds. By varying the size of some basis sets, we could demonstrate the validity of the repeatedly discussed dependency of the Mulliken MPA on the basis set.For the performance of the quantum mechanical calculations we employed the following methods and basis sets: HF/6-31G(d,p), DFT/B3PW91/6-31G(d), DFT/B3PW91/6-311++G(d,p), MP2/6-31G(d,p) and MP2/6-311++G(d,p).  相似文献   

2.
The structure of the diterpene alkaloid arcutin, which is isolated fromAconitum arcuatum Maxim. (Ranunculaceae), is solved by x-ray methods. Arcutin represents a new type of diterpene alkaloids containing a C5–C20 bond instead of the traditional C10–C20 bond in the carbon backbone.Translated from Khimiya Prirodnykh Soedinenii, No. 1, pp. 62–65, January–February, 2000.  相似文献   

3.
Thermal and infrared spectral studies of the urea-orthoboric acid complex are reported. The complex is formed through the elimination of 0.5 H2O molecule. Infrared spectral data show the presence of hydrogen-bonding and the force constant calculated for the N ... HO bond is found to be 4–5×10–5 dyne cm–1, which is for the order of a single bond and indicates strong hydrogen-bonding in the complex. A tentative structure for the complex is proposed.  相似文献   

4.
The electronic structure and stability of pyrrolyl are investigated using CASSCF, CASPT2 and G2(MP2) techniques. The ground state of pyrrolyl is found to be 2A2, with five π-electrons, as in cyclopentadienyl. The computed N–H bond energy of pyrrole is 94.8 kcal mol−1, while the heat of formation ΔfH298o of pyrrolyl is deduced to be 70.5±1 kcal mol−1. The Arrhenius parameters of N–H and C–H bond fission in pyrrole and cyclopentadiene and hydrogen abstraction reactions (by hydrogen) were also computed, indicating that pyrrolyl forms predominantly by C–H bond fission of pyrrolenine rather than by direct N–H bond fission.  相似文献   

5.
The kinetics of oxidation of l-methionine by chromic acid in acidic medium (pH=0.83–2.2) has been studied spectrophotometrically. The effect of l-methionine and chromium(VI) concentrations on the rate of the reaction was determined. The reaction rate decreases with increasing the pH of the medium. The kinetics of the formation of a chromium(III) complex conform to the rate law:with k1=7.5×10–2s–1 and Kes1=43.85 at constant [H+]=1.9×10–2moldm–3 and [l-methionine]T= 2.0×10–2moldm–3. The same values were found with [l-methionine]T variation at constant [H+]. The reaction proceeds through formation of chromium(VI)-l-methionine ester in a rapid pre-equilibrium step, followed by a slower redox reaction of the ester. The present study provides kinetic evidence for formation of a complex ion (ester). One mole of Cr2O72– oxidizes four moles of l-methionine, which acts as a monodentate ligand and binds to chromium(VI) through the sulfur atom. The coordinated sulfur atom of l-methionine–chromium(VI) ester is responsible for the oxidative degradation (breaking of the C1-C2 bond) of l-methionine. Coordinated oxygen of the carboxylate group inhibits the cleavage of the C1-C2 bond.  相似文献   

6.
Conclusions An x-ray study has been made on TGS at various temperatures above and below the transition point, which shows that the populations in the two positions for the disordered nitrogen atom in glycine 1 equalize as the temperature rises, which is accompanied by increases in the deviations of those positions from the C–CO2 mean-square plane along the polarization axis, which is accompanied by breakage of the bond between the NH3 + group in glycine 1 and the SO4 2–.Below the transition point, the hydrogen atom in the O(21)–H...O(32) bond is localized near the oxygen atom in glycine 2. Above that point, it is distributed between the two positions.The transition is of mixed type, as it has features of displacement-type and order-disorder transitions.Lomonosov University, Moscow. Translated from Zhurnal Strukturnoi Khimii, Vol., 31, No. 3, pp. 84–90, May–June, 1990.  相似文献   

7.
Density functional theory (DFT) using the finite cluster approach is utilized to compute binding energies, bond geometries, and vibrational properties of carbon monoxide adsorbed on Pt(111) as a function of the external interfacial field, focusing attention on the metal–CO bond itself. Comparison with electrode potential-dependent frequencies for the metal–CO (νM–CO) as well as the much-studied intramolecular C---O (νCO) vibration, as measured by in-situ Raman and infrared spectroscopy, facilitate their interpretation in terms of metal-chemisorbate bonding for this archetypal electrochemical system. Decomposing the calculated metal–CO binding energy and vibrational frequencies into individual orbital and steric repulsion components enables the role of such quantum-chemical interactions to the field- (and hence potential-) dependent bonding to be assessed. No simple relationship between the field(F)-dependent binding energies and the νM–CO frequencies is evident. While the DFT νM–COF slopes are negative at positive and small–moderate negative fields, reflecting the prevailing influence of back-donation, a νM–COF maximum is obtained at larger negative fields for atop CO, and a plateau for hollow-site CO. This Stark-tuning behavior reflects largely offsetting field-dependent contributions from π and σ surface bonding, and can also be rationalized on the basis of changes in the electrostatic component of νM–CO from increasing M–CO charge polarization. A rough correlation between the field-dependent νM–CO frequencies and the corresponding bond distances, rM–CO, is observed for hollow and atop CO in that rM–CO shortens towards less positive fields, but becomes near-constant at moderate–large negative fields. A more quantitative correlation between the field-dependent C---O frequencies and bond lengths is also evident. In harmony with earlier findings (and unlike the νM–COF behavior), the νCOF dependence is due chiefly to changes in the back-donation bonding component. The overall vibrational frequency-field behavior predicted by DFT is also in semi-quantitative concordance with experimental potential-dependent spectra.  相似文献   

8.
The reaction of [Cp′Cr(CO)2(μ-SBu)]2 (1) (Cp′ = MeC5H4) with (PPh3)2Pt(PhCCPh) gives Cp′Cr(CO)2(μ-SBu)Pt(PPh3)2 (2) which could be regarded as a product of the substitution of acetylene ligand at platinum by a monomeric chromium–thiolate fragment. According to the X-ray diffraction analysis 2 contains single Cr–Pt (2.7538(15)) and Pt–S (2.294(2) Å) bonds while Cr–S bond (2.274(3) Å) is shortened in comparison with ordinary Cr–S bonds (2.4107(4)–2.4311(4) Å) in 1. The bonding between Cr–S fragment and platinum atom is similar to the olefine coordination in their platinum complexes.  相似文献   

9.
Activation energies have been measured for the thermal reactions in the formation and bleaching of colored forms of spiropyranes containing various heterocyclics as liquids, in the vitreous amorphous state, and in solution in benzophenone. The activation energies have been compared for thermal coloring, as have the quantum yields in the photocoloring, which are related to the spiropyrane structures. The thermal breakage of the Csp –O bond in a spiropyrane produces an unstable Bx isomer of the colored form. Nitrospiropyranes and spironaphthopyranes then give rise to the cis-cisoid isomer of the colored form produced in the photochemical reaction, which has an absorption spectrum close to that of the relatively stable colored form. The activation energy for the formation of the Bx form is usually correlated with the Csp –O bond length in the ground state, in contrast to the quantum yield.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 23, No. 4, pp. 476–480, July–August 1987.  相似文献   

10.
Summary The structure of hydrofluorides with the composition MeF2 · HF(Me=Sr, Ba) includes two anions: F and HF2 with a linear hydrogen bond. In SrF2 · HF, the hydrogen bond is characterized by a more asymmetrical position of the proton than in BaF2 · HF.Increasing the number of added HF molecules leads to the formation of two bifluoride ions in the compounds CaF2 · 2HF and SrF2 · 2.5HF. The hydrogen bond in this case is also nonsymmetrical, and the average F–H distance is equal to 1.17 A. Thus, the hydrogen bond in these compounds is of the same order as in KF · 2HF (rF–F=2.33 A) [12],i.e., weaker than in KF · HF (rF–F=2.26 A) [13], but stronger than in solid HF (rF–F=2.49 A) [14].A structure with two polyfluoride anions evidently corresponds to the compound BaF2 · 3HF.Institute of Inorganic Chemistry, Siberian Branch of the Academy of Sciences of the USSR. Translated from Zhurnal Strukturnoi Khimii, Vol. 9, No. 2, pp. 202–206, March–April, 1968.  相似文献   

11.
Phenolic 2-arylcoumarans 16 were used to examine the behaviors of β-5 subunits in lignin during tetramethylammonium hydroxide (TMAH) thermochemolysis. Products were monitored by gas chromatography/mass spectrometry. The process predominantly provided dimeric products with the opened hydrofuran ring. Substituent changes at the γ-position of ring A and at the 5-position of ring B had a large effect on the product compositions. 2-Arylcoumarans 1 and 6 with the γ-CH2OH substituent predominantly gave 2,3,3′,4′-tetramethoxystilbenes involving the elimination of the γ-CH2OH substituent, while 25 with the γ-CH3 substituent gave a mixture of 2,3,3′,4′-tetramethoxy-α-methylstilbenes and α-methoxy-α-(3′,4′-dimethoxyphenyl)-β-(2,3-dimethoxyphenyl)propanes. Substituent –CHCHCH3 on ring B remained unaffected. Substituents –CHCHCH2OH and –COOH on ring B produced the corresponding methyl ether and ester, respectively, by methylation. The –CHCHCHO substituent on ring B was converted to the –CHO substituent.  相似文献   

12.
We present semiempirical quantum chemical calculations of geometric structures, charge distributions, energy levels, ionization potentials, and enthalpies of formation for aniline tetramers in different oxidation states using the semiempirical AM1 method. For tetraaniline radical cation the effect of these three counterions on the above-mentioned physicochemical parameters was analyzed. The ions studied included Cl, HSQ4 , and CH3COO. Chloride counterion showed a large charge transfer to the chain, as was shown in a preceding work [4]. HSO4 showed a strong charge stabilization without transfer. CH3COO exhibits hydrogen bond formation and also displays a strong charge stabilization in a fragment of the chain, but does not transfer any charge. Also, Cl was able to form a hydrogen bond, depending on the initial position that it occupies in relation to the tetrameric chain, which is then optimized. Charge transfer is present for the Cl cases. For dication structures the effect of SO 4 2– was analyzed. Our calculations showed that the distribution of the energy levels of tetraaniline radical cation can be comparable to the polyaniline ones by adding a counterion to the tetramer.  相似文献   

13.
The pure rotational spectrum of NaCH3 and NaCD3 in their states has been recorded using millimeter/sub-mm direct absorption techniques in the 300–510 GHz range. This work is the first gas-phase detection of sodium monomethyl, which was created by the reaction of sodium vapor with tetramethyl tin. Ten rotational transitions were measured for NaCH3 for the K=0 through K=5 components and, in select cases, up to K=10, and four transitions (K=0–7) for NaCD3. Rotational constants have been accurately determined for both isotopomers, suggesting a sodium–carbon bond length of 2.30 Å and an H–C–H bond angle of 107.3°.  相似文献   

14.
The purpose of this report is to quantitatively find the cause for the elongation of the R-C bond in R-COO (R = H, CH3 and C2H5) and the shortening of the C-O bond in CH3-O upon deprotonation in the gas phase. These elongations and shortenings result from the contributions of R---CO2 and H---CH2=O as resonance structures to the systems. Because these structures must make only a small contribution in the crystal, the R-C bond lengths of R-COO (R= H and CH3) in the crystal structure are shorter than those in the gas phase.  相似文献   

15.
The C–HX (X=N, O, S) intramolecular hydrogen bond between the α-hydrogen of the vinyl group and the corresponding heteroatom in the series of 1-vinyl-2-(2′-heteroaryl)pyrroles was examined by ab initio calculations at the B3LYP/6-311(d,p) level. It was shown that the C–HN hydrogen bond is stronger than the C–HO hydrogen bond and the latter is, in turn, stronger than the C–HS hydrogen bond. This conclusion is supported by calculations of 1H NMR chemical shieldings.  相似文献   

16.
An x-ray structural investigation of four permethylcyclocarbosilanes, whose heterocycles are formed by –Si–CH2–CH2–, –Si–CH=CH–, and –Si–CC– fragments, has been carried out. Strong distortions of some bond lengths and bond angles (caused by the shortness of the Si...Si distances between the neighboring Si atoms and the absence of intermolecular contancts with the participation of atoms of the heterocycles), which attest to statistical disordering of the molecules in the crystals, whose character could be established unequivocally for three of the compounds, have been observed in all the structures.For part LXIII, see [1].A. N. Nesmeyanov Institute of Organometallic Compounds, Academy of Sciences of the USSR. Translated from Zhurnal Strukturnoi Khimii, Vol. 32, No. 2, pp. 116–124, March–April, 1991.  相似文献   

17.
The chloride complexes of copper(II) (catalysts or catalyst precursors for various reactions of halogenated hydrocarbons) were characterized using electron, EPR, and EXAFS spectroscopy. It was found that chlorocuprates occur as mononuclear ([CuCl4]2–), binuclear ([Cu2Cl6]2–), and, probably, polynuclear species in chlorobenzene solutions. The Cu–Cl bond length in [CuCl4]2– is 2.25 ± 0.2 Å, which is close to the same values for crystalline tetrachlorocuprates. It was assumed that the chloride complexes of copper with counterions occur as globules in chlorobenzene.  相似文献   

18.
The homolytic dissociation of a single bond involves the decorrelation of one electron pair. Thus, the contribution of electron correlation to dissociation energies is large. In the present paper a new procedure is presented which allows the computation of the (within the given basis) complete correlation energy of one optimized electron pair. The method which requires only modest computational effort has been applied to the calculation of dissociation energies of a number of bonds of different types. The results show that the correlation of the electron pair of the bond which is broken contributes about 50–80% to the change of the total correlation energy occuring during the dissociation process which amounts to 20–70 kcal/mol. The fraction of correlation contributed by the bond electron pair as well as the relative importance of the left-right correlation within the bond depend very much on the type of the bond. In the case of CC and CH single bonds our method yields dissociation energies which are low by only about 5 kcal/mol. Thus, the method seems to be well suited for the calculation of potential surfaces of non-concerted organic chemical reactions which involve diradicals as intermediates.  相似文献   

19.
The C–HN hydrogen bond in the methane–ammonia complex is studied by determining its bond dissociation energy (BDE) and the n(N)→σ*(C–H) interaction. At the MP2(Full)/6-311++G(3df,2p) level of theory with basis set superposition error (BSSE) correction, the BDE was determined to be 2.5 kJ mol−1. The n(N)→σ*(C–H) interaction at this level of theory was found to be 3.7 kJ mol−1 by natural bond orbital (NBO) analysis. It was also found that the NBO values are in general higher than the BDE values with BSSE correction when they are compared at the same level of theory.  相似文献   

20.
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   

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