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 15 毫秒
1.
In earlier hydroformylation studies modification of the rhodium catalyst with o-methyl-substituted or o-ethyl-substituted phosphane ligands have increased regioselectivity to branched aldehydes. The promising results achieved created a need for further studies. Hence, a wider group of o-substituted arylphosphane ligands, e.g. (2-cyclohexylphenyl)diphenylphosphane, (2-isopropylphenyl)diphenylphosphane, (2-methylnaphthyl)diphenylphosphane, (2,5-dimethylphenyl)diphenylphosphane and (2-phenylphenyl)diphenylphosphane were synthesized and tested in rhodium-catalyzed hydroformylation to support the previous findings. Characterization of the ligands was made by NMR spectroscopy (1H, 31P{1H}, 13C{1H}, HSQC, COSY-90 and COLOC). Additional parameters for evaluation of the stereoelectronic properties of the ligands were provided by quantum mechanical calculations and by synthesizing Rh(acac)(CO)(PR3) complexes. In the rhodium-catalyzed hydroformylation of propene and 1-hexene the ligands increased the formation of branched aldehydes compared to triphenylphosphane. Additionally the increasing size of the o-alkyl-substituent was found to effect favorably to the iso-selectivity.  相似文献   

2.
The synthesis of dimethyl {2-[3-(4-nitrophenyl)-1H-pyrazol-5-yl]ethyl}malonate monohydrate 1, C16H17N3O6·H2O was performed and the molecular structure has been studied by using NMR, single crystal X-ray diffraction and ab initio calculations. The title compound presents a pyrazole ring (N1 to C5), a phenyl ring (C1″ to C6″) attached to C3 and the ethylene dimethyl malonate frame (C1′ to C7′) attached to C5. The torsion angle defined by N2C3C1″C2″ (−12.26°) showed that pyrazole and phenyl rings are not in the same plane. Monohydration in (1) is present in the structure by a NHOH2 hydrogen bonding, with a bond length of 1.782 Å. Experimental and theoretical evidences indicated the preference of the 3-tautomer over the corresponding 5-tautomer in the titled pyrazole.  相似文献   

3.
Quantum chemical methods (MP2 and B3LYP) together with a topological analysis of the charge density have been used to study the BH3- or BF3-mediated reaction of benzonitrile oxide with acetonitrile, propyne and propene. In the reaction with propene or propyne, addition of Lewis acids has only little influence on the outcome of the reactions. The cycloaddition of nitrile oxides with nitriles, however, is generally promoted by strong Lewis acids. When the Lewis acid coordination takes place at the nitrile oxide the reactant is activated and the product binds weakly to the Lewis acid so that the reaction is expected to be catalytic. In the case of coordination to the nitrile the reaction is Lewis acid mediated. Here the reactant is not much influenced by addition of Lewis acid, but the transition state and the product are stabilised and consequently such processes require a stoichiometric amount of Lewis acid and form a stable Lewis acid-product complex.It has also been demonstrated that the different activation routes for these reactions involve different reaction mechanisms. Whereas the reaction of a Lewis acid coordinated nitrile oxide is of ‘inverse electron demand’, the Lewis acid coordinated nitrile reacts through a ‘normal electron demand’ cycloaddition.  相似文献   

4.
A novel cycloadduct of 1-boryl-3,4-dimethylphosphole was prepared by reaction of 3,4-dimethylphospholyl anion with monobromoborane-methylsulfide complex (CH3)2S · BH2Br at −60 °C. It was characterized as a six-membered trimer by spectroscopic means, and its structure confirmed by an X-ray crystal analysis and quantum chemical calculations. Density functional theory calculations (B3LYP) showed that the cyclic trimer is by far more stable than the monomer, dimers or open-chain forms. Various molecular and spectroscopic properties of the borylphosphole monomer and trimer were evaluated. In particular, the changes of the 31P NMR chemical shifts upon oligomerization were examined. The six-membered ring was demonstrated to exist preferentially in a chair-like conformation. Computed NMR chemical shifts (1H, 13C and a lesser extent 31P) appear to be a highly sensitive analytical tool for distinguishing ring conformations having only small energy differences.  相似文献   

5.
Structures and energies of cyclo-C5H5SF and cyclo-C5H5SF3 have been calculated. In both cases the 2- and 4-CF-isomers are more stable than the SF and SF3 isomers. The fluxional behavior of the sulfur bonded fluorides has been calculated also. In cyclo-C5H5SF an ellipsoidal rotation of the sulfur bonded fluorine atom is observed with a barrier of a few kcal mol−1. In sulfur bonded cyclo-C5H5SF3 the (Turnstile) rotation is predicted to occur without noticeable barrier, in agreement with previous work.Attempts to isolate the sulfur bonded isomers failed entirely: always 2 or 4-carbon-fluorides were obtained for cyclo-C5H5SF. The acyclic SF5 carrying precursors for the synthesis of cyclo-C5H5SF3 failed in crucial steps of the reactions.  相似文献   

6.
The details of weak C–Hπ interactions that control several inter and intramolecular structures have been studied experimentally and theoretically for the 1:1 C2H2–CHCl3 adduct. The adduct was generated by depositing acetylene and chloroform in an argon matrix and a 1:1 complex of these species was identified using infrared spectroscopy. Formation of the adduct was evidenced by shifts in the vibrational frequencies compared to C2H2 and CHCl3 species. The molecular structure, vibrational frequencies and stabilization energies of the complex were predicted at the MP2/6-311+G(d,p) and B3LYP/6-311+G(d,p) levels. Both the computational and experimental data indicate that the C2H2–CHCl3 complex has a weak hydrogen bond involving a C–Hπ interaction, where the C2H2 acts as a proton acceptor and the CHCl3 as the proton donor. In addition, there also appears to be a secondary interaction between one of the chlorine atoms of CHCl3 and a hydrogen in C2H2. The combination of the C–Hπ interaction and the secondary ClH interaction determines the structure and the energetics of the C2H2–CHCl3 complex. In addition to the vibrational assignments for the C2H2–CHCl3 complex we have also observed and assigned features owing to the proton accepting C2H2 submolecule in the acetylene dimer.  相似文献   

7.
Novel calix[4]arene derivatives functionalized with two or four isosteviol units at the upper rim and a new calix[6]arene having six isosteviol moieties at the lower rim have been synthesized. The structures of these compounds have been confirmed by NMR and mass spectrometry data. All 1H and 13C NMR chemical shifts of isosteviol were fully assigned by extensive NMR spectroscopic methods, and used to clarify the structures and conformations of isosteviol-calixarene conjugates.  相似文献   

8.
Increasingly pronounced hydrocarbon character is exhibited by C6H6B12, the first unsubstituted hexacarbaborane, and CH3C5B7H12, the first cluster pentacarbaborane. These compounds shed light on the structural dichotomy between open hydrocarbon skeletons and polyhedral borane frameworks for high-carbon carboranes.  相似文献   

9.
Since the discovery of the first drum-like CoB16- complex, metal-doped drum-like boron nanotubular structures have been investigated with various metal dopants and different tubular size, forming a new class of novel nanostructures. The CoB16- cluster was found to be composed of a central Co atom coordinated by two fused B8 rings in a tubular structure, representing the potential embryo of metal-filled boron nanotubes and providing opportunities to design one-dimensional metal-boron nanostructures. Here we report improved photoelectron spectroscopy and a more in-depth electronic structure analysis of CoB16-, providing further insight into the chemical bonding and stability of the drum-like doped boron tubular structures. Most interestingly, we find that the central Co atom has an unusually low oxidation state of ?1 and neutral CoB16 can be viewed as a charge transfer complex (Co-@BB16+), suggesting both covalent and electrostatic interactions between the dopant and the boron drum.  相似文献   

10.
Diffuse reflectance Fourier transform infrared (DRIFT) spectroscopy was used to study the adsorption process of the water-soluble polyacrylic acid (PAA) polymer on hydrous δ-Al2O3. Vibrational assignment of PAA, sodium polyacrylate, (Na–PA) and the PA-oxide surface complex was achieved by comparison of observed band position and intensity in the DRIFT spectra with wavenumbers and intensities from ab initio quantum mechanical calculations. The presented data of polyacrylic acid suggest that IR data calculated ab initio on relatively short oligomers (quantum-mechanical oligomer approach) may provide valuable information regarding the interpretation of polyelectrolyte infrared spectra. Batch adsorption experiments were performed to sorb PAA onto the δ-Al2O3 surface. The results obtained from DRIFT studies were compared with adsorption isotherm experiments in order to relate the level of PAA coverage to the nature of the surface complex. Ab initio molecular orbital calculations on PAA/Al2O3 clusters were used to model possible surface complexes. Strong correlation were found between theoretical and observed DRIFT frequencies of the antisymmetric R-COO vibration. A number of possible configurations of the polyacrylic acid/aluminate surface complex were tested via ab initio calculations. These possible configurations included different di-aluminium octahedral Al3+ surface models. Results obtained from adsorption isotherm experiments, DRIFT spectra and ab initio calculations indicate that the carboxylate oxygens bridge an Al3+-octahedral dimer [Al2(OH)24(H2O)2(OH)] in a ligand-exchange inner sphere complex.  相似文献   

11.
Hydroquinone (HYQ) in the lowest electronically excited state has been studied by ab initio quantum chemical calculations and resonant two-photon ionization (R2PI) spectroscopy. Calculations at the MP2/6-31G* and CIS/6-31G* levels yield satisfactory results on structures and vibrational frequencies of the cis-HYQ and trans-HYQ in both the S0 and S1 states. Only transitions involving in-plane modes are observed in the R2PI spectrum of HYQ. All spectral bands including some newly observed ones have been successfully assigned with the help of our computed results and analogy with the reported spectra for similar molecules.  相似文献   

12.
13.
The molecular structure of phenylsilane has been determined accurately by gas-phase electron diffraction and ab initio MO calculations at the MP2(f.c.)/6-31G* level. The calculations indicate that the perpendicular conformation of the molecule, with a Si–H bond in a plane orthogonal to the plane of the benzene ring, is the potential energy minimum. The coplanar conformation, with a Si–H bond in the plane of the ring, corresponds to a rotational transition state. However, the difference in energy is very small, 0.13 kJ mol−1, implying free rotation of the substituent at the temperature of the electron diffraction experiment (301 K). Important bond lengths from electron diffraction are: <rg(C–C)>=1.403±0.003 Å, rg(Si–C)=1.870±0.004 Å, and rg(Si–H)=1.497±0.007 Å. The calculations indicate that the Cipso–Cortho bonds are 0.010 Å longer than the other C–C bonds. The internal ring angle at the ipso position is 118.1±0.2° from electron diffraction and 118.0° from calculations. This confirms the more than 40-year old suggestion of a possible angular deformation of the ring in phenylsilane, in an early electron diffraction study by F.A. Keidel, S.H. Bauer, J. Chem. Phys. 25 (1956) 1218.  相似文献   

14.
(n)J(Se,Se) (n=1-4) nuclear couplings between Se atoms were analyzed by using molecular orbital (MO) theory as the first step to investigating the nature of bonded and nonbonded (n)J(Se,Se) interactions between Se atoms. The values were calculated by employing Slater-type triple xi basis sets at the DFT level, which were applied to structures optimized with the Gaussian 03 program. The contribution from each occupied MO (psi(i)) and psi(i)-->psi(a) (psi(a)=unoccupied MO) transition was evaluated separately. 1J(Se,Se) was calculated for the MeSeSeMe model compound, which showed a typical dependence on the torsion angle (phi(C(Me)SeSeC(Me))). This dependence explains the small values (< or =64 Hz) of 1Jobsd(Se,Se) observed for RSeSeR' and large values (330-380 Hz) of 1Jobsd(Se,Se) observed for 4-substituted naphtho[1,8-c,d]-1,2-diselenoles, which correspond to synperiplanar diselenides. The HOMO-->LUMO and HOMO-1-->LUMO transitions contribute the most to 1J(Se,Se) at phi=0 and 180 degrees to give large values of 1J(Se,Se), whereas various transitions contribute and cancel each other out at phi=90 degrees to give small values of 1J(Se,Se). Large 4Jobsd(Se,Se) values were also observed in the nonbonded Se...Se, Se...Se=O, and O=Se...Se=O interactions at naphthalene 1,8-positions. The Fermi contact (FC) term contributes significantly to 4J(Se,Se), whereas the paramagnetic spin-orbit (PSO) term contributes significantly to 1J(Se,Se). 2J(Se,Se) and 3J(Se,Se) were analyzed in a similar manner and a torsional angular dependence was confirmed for 3J(Se,Se). Depending on the structure, the main contribution to (n)J(Se,Se) (n=2, 3) is from the FC term, with a lesser contribution from the PSO term. Analysis of each transition enabled us to identify and clearly visualize the origin and mechanism of the couplings.  相似文献   

15.
Ab initio molecular orbital calculations are reported for small neutral molecules and cations containing magnesium, nitrogen and hydrogen. Structures have been optimized using gradient techniques at B3LYP/6-31+G(d) and at MP2(full)/6-311++G(d,p). Single-point calculations are reported at QCISD(T)(full)/6-311++G(2df,p) and at CCSD(T)(full)/6-311++G(2df,p) levels using geometries optimized at MP2(full)/6-311++G(d,p). Standard enthalpies of formation at 298 K have been calculated at these two higher levels of theory. Other thermochemical properties calculated include ionization energies and proton affinities. The binding enthalpies of ammonia to Mg+, MgNH2+ and MgNH3+ are also reported.  相似文献   

16.
The title compound of p-trifluoromethylphenyl thioacid amide has been synthesized in one step and characterized by elemental analysis, UV and X-ray single crystal diffraction. Ab initio calculations indicate that both HF/6-311G** and B3LYP/6-311G** methods can reproduce the title compound well. Electronic absorption spectra calculated by the time-dependent density functional theory (TD-DFT) show that the two absorption bands are mainly derived from the contribution of bands π → π*. Thermodynamic properties of the title compound have been predicted based on the optimized structure. The calculation of the second order optical nonlinearity also has been carried out, and the molecular hyperpolarizability is 2.31770 × 10−30 esu.  相似文献   

17.
Ligand effects on chain termination reactions in hafnocene-catalyzed ethene polymerization process have been systematically studied by quantum chemical methods. β-hydrogen transfer to metal, β-hydrogen transfer to monomer and hydrogenolysis were studied for 27 hafnocenes, initiating the chain termination reactions after insertion of the second ethene monomer. The results of the calculations were studied as a function of the ligand structure, focusing on the effects of various ancillary ligands, ligand substituents and bridging units. The ligand effects on chain termination reactions are strongly affected by combined effects of various structural units, in particular, in the cases of β-hydrogen transfer to monomer and hydrogenolysis. The results are expected to aid in design and development of new hafnocene polymerization catalysts.  相似文献   

18.
The heptasilane Me(SiMe3)2SiSiH2SiMe(SiMe3)2 was synthesized from Me(SiMe3)2SiK and H2Si(OSO2CF3)2. Crystals suitable for a X-ray single crystal analysis could be grown, with the somewhat surprising result that the two dihedral angles (H3)CSiSi(H2)Si are different in the crystal (24.58(10)° and 31.67(11)°). SiSiSi-bonds angles are widened, with values up to 117°. Ab initio calculations at the density functional B3LYP level employing 6-311G(d) basis sets predict minima for five conformers 1-5 with relative energies 0.0, 3.1, 8.2, 10.8 and 18.1 kJ/mol, respectively. Moreover, SiSiSiSi dihedral angles spanning the range 43.5-172.3° are predicted, reflecting the small forces which are required for distorting these angles.In the Raman spectrum of a solution in toluene, three lines at 350, 340 and 330 cm−1 are observed in a wavenumber range which is typical for the SiSi-pulsation of methylated oligosilanes. The relative intensity ratio of the bands is temperature dependent, reflecting the changes in conformer concentrations that occur according to Boltzman’s law. Supported by the ab initio calculations, the Raman band at 350 cm−1 is assigned to an ‘averaged’ conformer 1 and 2, because a rapid interconversion between 1 and 2 has to be assumed due to a small barrier separating them. The bands with wavenumbers 340 and 330 cm−1 originate from conformers 3 and 4. From the Raman spectra, relative energies 0.0 (1 + 2), 2.2 (3) and 6.3 (4) kJ/mol are deduced, the presence of 5 is not observed. Caused by solvent effects, these values differ somewhat from the ab initio results.  相似文献   

19.
Plain rules founded in a theoretical background are presented that can be used to determine the structure of selenium compounds on the basis of delta(Se) data and to predict delta(Se) data from a given structure with satisfactory accuracy. As a first step to establish such rules, the origin of delta(Se) is elucidated on the basis of MO theory. The Se(2-) ion was chosen as the standard for the analysis. The concept of the pre-alpha effect is proposed, which is defined as the downfield shift due to protonation of a lone-pair orbital of Se. The pre-alpha effect of two protons in H(2)Se is explained by the generation of double sigma(Se--H) and sigma*(Se--H) through protonation of the spherical Se(2-) ion. The orbitals, together with n(p)(Se), result in effective transitions for the pre-alpha effect. The alpha effect is the downfield shift caused by the replacement of Se--H by Se--Me. The extension of HOMO-2 [4p(y)(Se)], HOMO-1 [4p(x)(Se)], and HOMO [4p(z)(Se)] over the whole Me(2)Se molecule is mainly responsible for the alpha effect. The beta effect originates not from the occupied-to-unoccupied (psi(i)-->psi(a)) transitions but from the occupied-to-occupied (psi(i)-->psi(j)) transitions. Although psi(i)-->psi(j) transitions contribute to upfield shifts in Me(2)Se, the magnitudes become smaller as the methyl protons are substituted by Me groups one after another. The gamma effect of upfield shifts is also analyzed, although complex. The effect of p(Se)-pi(C==C) conjugation is analyzed in relation to the orientational effect. Contributions from each MO (psi(i)) and each psi(i)-->psi(a) transition are evaluated separately, by using a utility program derived from the Gaussian 03 program suite (NMRANAL-NH03G). The treatment enables us to visualize and understand the origin of (77)Se NMR chemical shifts.  相似文献   

20.
Two different transition structures (TSs) have been located and characterized for the chorismate conversion to prephenate in Bacillus subtilis chorismate mutase by means of hybrid quantum-mechanical/molecular-mechanical (QM/MM) calculations. GRACE software, combined with an AM1/CHARMM24/TIP3P potential, has been used involving full gradient relaxation of the position of ca. 3300 atoms. These TSs have been connected with their respective reactants and products by the intrinsic reaction coordinate (IRC) procedure carried out in the presence of the protein environment, thus obtaining for the first time a realistic enzymatic reaction path for this reaction. Similar QM/MM computational schemes have been applied to study the chemical reaction solvated by ca. 500 water molecules. Comparison of these results together with gas phase calculations has allowed understanding of the catalytic efficiency of the protein. The enzyme stabilizes one of the TSs (TSOHout) by means of specific hydrogen bond interactions, while the other TS (TSOHin) is the preferred one in vacuum and in water. The enzyme TS is effectively more polarized but less dissociative than the corresponding solvent and gas phase TSs. Electrostatic stabilization and an intramolecular charge-transfer process can explain this enzymatically induced change. Our theoretical results provide new information on an important enzymatic transformation and the key factors responsible for efficient selectivity are clarified. Received: 25 March 2000 / Accepted: 7 August 2000 / Published online: 23 November 2000  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