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 116 毫秒
1.
The molecular structure and benzene ring distortions of ethynylbenzene have been investigated by gas-phase electron diffraction and ab initio MO calculations at the HF/6-31G* and 6-3G** levels. Least-squares refinement of a model withC 2v, symmetry, with constraints from the MO calculations, yielded the following important bond distances and angles:r g(C i -C o )=1.407±0.003 Å,r g(C o -C m )=1.397±0.003 Å,r g(C m -C p )=1.400±0.003 Å,r g(Cr i -CCH)=1.436 ±0.004 Å,r g(C=C)=1.205±0.005 Å, C o -C i -C o =119.8±0.4°. The deformation of the benzene ring of ethynylbenzene given by the MO calculations, including o-Ci-Co=119.4°, is insensitive to the basis set used and agrees with that obtained by low-temperature X-ray crystallography for the phenylethynyl fragment, C6H5-CC-, in two different crystal environments. The partial substitution structure of ethynylbenzene from microwave spectroscopy is shown to be inaccurate in the ipso region of the benzene ring.  相似文献   

2.
Neutral η1-benzylnickel carbene complexes, [Ni(η1-CH2C6H5)(IiPr)(PMe3)(Cl)] (3) (IiPr = 1,3-bis-(2,6-diisopropylphenyl)imidazol-2-ylidene) and [Ni(η1-CH2C6H5)(SIiPr)(PMe3)(Cl)] (4) (SIiPr = 1,3-bis-(2,6-diisopropylphenyl)imidazolin-2-ylidene), were prepared by the reaction between [Ni(η3-CH2C6H5)(PMe3)(Cl)] and an equivalent amount of the corresponding free N-heterocyclic carbene. The preparation of η3-benzylnickel carbene complexes, [Ni(η3-CH2C6H5)(IiPr)(Cl)] (5) and [Ni(η3-CH2C6H5)(SIiPr)(Cl)] (6) were carried out by the abstraction of PMe3 from 3 and 4 by the treatment of B(C6F5)3. The treatment of AgX on 5 and 6 produced the anion-exchanged complexes, [Ni(η3-CH2C6H5)(NHC)(X)] (7, NHC = IiPr, X = O2CCF3; 8, NHC = IiPr, X = O3SCF3; 9, NHC = SIiPr, X = O2CCF3; 10, NHC = SIiPr, X = O3SCF3). The solid state structures of 3 and 10 were determined by X-ray crystallography. The η3-benzyl complexes of IiPr (5, 7, and 8) alone, in the absence of any activators such as borate and MAO, showed good catalytic activity towards the vinyl-type norbornene polymerization. The catalyst was thermally robust and the activity increases as the temperature rises to 130 °C.  相似文献   

3.
Drastic effects of Lewis acids E(C6F5)3 (E = Al, B) on polymerization of functionalized alkenes such as methyl methacrylate (MMA) and N,N-dimethyl acrylamide (DMAA) mediated by metallocene and lithium ester enolates, Cp2Zr[OC(OiPr)CMe2]2 (1) and Me2CC(OiPr)OLi, are documented as well as elucidated. In the case of metallocene bis(ester enolate) 1, when combined with 2 equiv. of Al(C6F5)3, it effects highly active ion-pairing polymerization of MMA and DMAA; the living nature of this polymerization system allows for the synthesis of well-defined diblock and triblock copolymers of MMA with longer-chain alkyl methacrylates. In sharp contrast, the 1/2B(C6F5)3 combination exhibits low to negligible polymerization activity due to the formation of ineffective adduct Cp2Zr[OC(OiPr)CMe2]+[OC(OiPr)CMe2B(C6F5)3] (2). Such a profound Al vs. B Lewis acid effect has also been observed for the lithium ester enolate; while the Me2CC(OiPr)OLi/2Al(C6F5)3 system is highly active for MMA polymerization, the seemingly analogous Me2CC(OiPr)OLi/2B(C6F5)3 system is inactive. Structure analyses of the resulting lithium enolaluminate and enolborate adducts, Li+[Me2CC(OiPr)OAl(C6F5)3] (3) and Li+[Me2CC(OiPr)OB(C6F5)3] (4), coupled with polymerization studies, show that the remarkable differences observed for Al vs. B are due to the inability of the lithium enolborate/borane pair to effect the bimolecular, activated-monomer anionic polymerization as does the lithium enolaluminate/alane pair.  相似文献   

4.
Co(II) complexes with 4,6-di(tert-butyl)-2-aminophenol (HLI) and 2-anilino-4,6-di(tert-butyl)phenol (HLII) have been synthesized and characterized by means of physico-chemical methods. The compounds HLI and HLII coordinate in their singly deprotonated forms and behave as bidentate O,N-coordinated ligands; their low-spin Co(II) complexes are characterized by CoN2O2 coordination modes and square planar geometry. Both the free ligands and their Co(II) and Cu(II) complexes (we have produced and characterized the latter before) exhibit a pronounced antifungal activity against Aspergillus niger, Fusarium spp., Mucor spp., Penicillium lividum, Botrytis cinerea, Alternaria alternata, Sclerotinia sclerotiorum, Monilia spp., which in a number of cases is comparable with that of Nystatin and Terbinafine or even higher. The reducing properties of the ligands and their metal(II) complexes, as well as their antifungal activities, were found to decrease in the order: Cu(LI)2 > Cu(LII)2 ? Co(LI)2 > Co(LII)2 > HLI > HLII.  相似文献   

5.
The influence of Cu(II) impurity on chemical equilibria in unsaturated and saturated ammonium oxalate (AO) aqueous solutions was investigated as a function of concentration cici of impurity. Using the computer programme “Hyss” the species present in the solutions were analysed. It was found that in the aqueous solutions of ammonium oxalate containing Cu(II) ions the following species are formed: Cu2+, Cu(OH)+, Cu(OH)2, CuC2O40 and Cu(C2O4)22− in addition to C2O42−, HC2O4, H2C2O4 and (NH4)2C2O40 species, and their concentration depends on concentrations cici of Cu(II) impurity and c of ammonium oxalate. The dependences of solution pH and of absorbance A   and the corresponding wavelength λλ for unsaturated aqueous solutions on ammonium oxalate concentration c   containing different concentrations cici of Cu(II) ions showed three well-defined regions characterised by transition values of solution pH or solute concentration c. Speciation analysis revealed that Cu2+ and CuC2O40, CuC2O40 and Cu(C2O4)22−, and Cu(C2O4)22− complexes are predominantly present in the solute concentration intervals c≤0.01c0.01 mol/dm3, 0.01 mol/dm3 <c<0.03<c<0.03 mol/dm3 and c≥0.03c0.03 mol/dm3, respectively. The concentration interval range 0.01 mol/dm3 <c<0.03<c<0.03 mol/dm3 corresponds to the pH interval where Cu(OH)2 is precipitated. It was found that the solubility of ammonium oxalate at 30 °°C increases practically linearly with an increase in the concentration of Cu(II) impurity. Speciation analysis of saturated aqueous solutions of ammonium oxalate revealed that Cu(II) ions contained in AO saturated solutions exist mainly as Cu(C2O4)22−-type complexes, and the increase in the solubility of AO in the presence of Cu(II) impurity is essentially due to an increase in the ratio of the concentrations of CuC2O40 and Cu(C2O4)22− species.  相似文献   

6.
The (liquid + liquid) solubility curves have been determined by a synthetic method for six binary mixtures of [acetonitrile + {heptyl methyl ether CH3OnC7H15, or ethyl hexyl ether C2H5OnC6H13, or pentyl propyl ether nC3H7OnC5H11, or isopentyl propyl ether nC3H7Oi C5H11, or dibutyl ether nC4H9OnC4H9, or butyl isobutyl ether nC4H9OiC4H9}]. The possibility of the COSMO-SAC model to account for the thermodynamic differences between these systems has been tested and the discussion on the influence of screening charge of ethers on the system properties was undertaken.  相似文献   

7.
A synthesis of N-acetylcolchinol, a key intermediate in the synthesis of ZD6126, was developed. The enantiodifferentiating step required the catalytic asymmetric hydrogenation of an enamide. After screening a range of metal and ligand combinations it was found that (S,S)-iPr-FerroTANE Ru(methallyl)2 and [(S,S)-tBuFerroTANE Rh(COD)]BF4 gave both high enantioselectivity (>90% ee) and high catalyst utility (molar S/C = 1000).  相似文献   

8.
The reactivity of N1-alkylsulfonyl- and N1-arylsulfonyl-2′,3′,5′-tri-O-acetylinosine with benzylamine and with 15NH3, regarding the attack on C2, has been shown to be in the order CF3SO2 (Tf) > 2,4-(NO2)2C6H3SO2 (DNs) ? 4-NO2C6H4SO2 (pNs) ≈ C6F5SO2 (PFBs) > 2-NO2C6H4SO2 (Ns) ? CH3SO2 (Ms) > 4-CH3C6H4SO2 (Ts) > 2,4,6-(CH3)3C6H2SO2 (Mts). In spite of its intermediate reactivity, the Ns group is the most appropriate, since in this case the formation of by-products is minimised during the ring-opening and ring-closing steps of the process. Another advantage of the Ns group is thus disclosed.  相似文献   

9.
The oligomerization and/or polymerization of ethylene catalyzed by the cationic η3-benzylcomplexes [Ni(η3-CH2C6H4-p-CF3)(P-P)]+ BPh4 (P-P=iPr2P(CH2)nPiPr2, n=1-3) have been studied. The activity of these single component catalysts depends on the length of the (CH2)n bridge of the diphosphine ligand. Thus, the dippm derivative (n=1) displays higher activity than compounds of the dippe (n=2) or dippp (n=3) ligands. The molecular weight of the products is also a function of n, and varies in the order dippm > dippe > dippp, with the former two catalysts giving rise to low molecular weight polyethylenes and the latter to oligomers.  相似文献   

10.
Infrared emission has been recorded from a heated seeded supersonic primary beam of HCl or HF (1) prior to collision with a target beam, and (2) subsequent to that collision. Mean collision energy and collision partner were varied systematically. After correction for elastic scattering, the net population change due to inelastic scattering in a translation—rotation (T ? R) energy-transfer encounter was obtained for specific J states ranging from J = 0–16 of vibrational level υ = 1 of the primary-beam molecule. The broad picture is that a net transfer into low-J states out of higher-J states takes place at low collision energies, and the converse at high collision energies. These observations are interpreted in terms of the “exponential model” for the relative cross sections of T ? R inelastic collisions, SR (JiJf), proposed earlier [J.C. Polanyi and K.B. Woodall, J. Chem. Phys. 56 (1972) 1563], modified here to satisfy microscopic reversibility. The constant C in the model, which governs the exponential decrease in SR with increasing energy difference ΔEJ between Jf and Ji, can be derived, as a function of collision energy T, from the present experimental data; C decreases as T increases, i.e. larger ΔJ become more probable. In order to check the validity of the model, it was compared with 3D trajectory results; according to this criterion it was found to give a very good representation of SR(JiJf) with a single value for C, within a limited range of Ji. The collision partners HCl + HF exhibit anomalously efficient rotational deactivation; evidence is presented which indicates that at low collision energies this is due to resonant R → R transfer. Very efficient deactivation of HCl by HCl, at low collision energy, is likely to be due to V — V transfer.  相似文献   

11.
Treatment of [M(H2Li)] with UCl4 in pyridine led to the formation of the dinuclear complexes [MLi(py)UCl2(py)2] and/or [Hpy][MLi(py)UCl3] [Li = N,N′-bis(3-hydroxysalicylidene)-R, R = 1,2-phenylenediamine (i = 1), R = trans-1,2-cyclohexanediamine (i = 2), R = 2-amino-benzylamine (i = 3), R = 1,3-propanediamine (i = 4), R = 2,2-dimethyl-1,3-propanediamine (i = 5); M = Cu or Ni]. The crystal structures show that the 3d and 5f ions occupy, respectively, the N2O2 and O4 cavities of the Schiff base ligand, the U4+ ion adopting a dodecahedral or pentagonal bipyramidal configuration in the neutral and anionic complexes, respectively.  相似文献   

12.
A series of chiral ansa-zirconocene ester enolate complexes incorporating C2- or Cs-symmetric ligands, including neutral rac-(EBI)ZrCl[OC(OiPr)CMe2] (1), rac-(EBI)Zr(OTf)[OC(OiPr)CMe2] (2), rac-(EBI)Zr(OTf)[OC(OMe)C(Me)CH2C(Me2)C(OiPr)O] (3), [Me2C(Cp)(Flu)]ZrMe[OC(OiPr)CMe2] (4), and cationic [Me2C(Cp)(Flu)]Zr+(THF)[OC(OiPr)CMe2][MeB(C6F5)3] (5), have been synthesized. Within the neutral C2-ligated zirconocene ester enolate series, the chloride derivative 1 is inactive toward any methyl methacrylate (MMA) additions, the methyl derivative rac-(EBI)ZrMe[OC(OiPr)CMe2] adds cleanly only 1 equiv. of MMA, and the triflate derivative 2 can add either 1 equiv. of MMA to form the single-MMA-addition product 3 or multiple equivalents of MMA to form P(MMA). Unlike the Cs-ligated methyl cation [Me2C(Cp)(Flu)ZrMe]+, which is inactive for MMA polymerization under various conditions, the Cs-ligated ester enolate cation 5 is moderately active for polymerization of MMA and N,N-dimethylacrylamide at ambient temperature; the resulting P(MMA) has a high molecular weight of Mn = 388 000 Da but a low syndiotacticity of [rr] = 64%, and the polymerization conforms to a chain-end control mechanism.  相似文献   

13.
RSeCCPh (1a, R = Et; 1b, R = n-Bu; 1c, R = Ph; 1d, R = 2,4,6-Me3C6H2) reacts with equimolar amounts of Fe2(CO)9 (2) to give [(μ-SeR)(μ-σ,π-CCPh)]Fe2(CO)6 (3a, R = Et; 3b, R = n-Bu; 3c, R = Ph; 3d, R = 2,4,6-Me3C6H2).Complexes 3a-3d exist as two isomers, depending on the axial or equatorial position of R at selenium.Addition of P(OiC3H7)3 (4) to 3d affords {(μ-Se-2,4,6-Me3C6H2)[μ-η1-CCPh(P(OiC3H7)3)]}Fe2(CO)6 (5) along with {(μ-Se-2,4,6-Me3C6H2)[μ-η11-PhCC(P(OiC3H7)3)]}Fe2(CO)6 (6).The solid-state structures of 3d, 5 and 6 were determined by single X-ray structure analysis.In mononuclear 3d the Fe(CO)3 fragments are bridged by a μ-Se-2,4,6-Me3C6H2 and a μ-σ,π-CCPh unit, resulting in an over-all butterfly arrangement.Due to steric reasons, the mesityl group is pointing away from the PhCC entity and hence, is located in an equatorial position.Compounds 5 and 6, which co-crystallise in the ratio of 7:93, feature aμ-bridging 2,4,6-Me3C6H2Se unit and either a vinylidenic CCPh(P(OiC3H7)3) (complex 5) or a olefinic PhCC(P(OiC3H7)3) (complex 6) building block of which the latter entity is part of a diiron cyclobutene ring.  相似文献   

14.
The impact of the HF cluster size on the proton-transfer switch between N?H-F and N-H?F in the anilide-(HF)n = 1-4 complexes was investigated by means of the quantum chemical methods. The change in the H-bond strength due to variation of the HF cluster size was well monitored by change in the binding energy (BE), structural parameter, electron density topology, natural charge and charge transfer. For n = 1, our results at the MP2/6-311++G(2d,2p) level show that the minimum-energy structure corresponds to the H-bonded complex PhNH?HF with excess negative charge localized on the N atom of the anilide anion. For n > 1, minimum energy structures correspond to PhNH2?F(HF)1-3 ones, namely a solvated F ion. This is a case in which the relative change in the acidity of the HF is observed in the ground state as the size of cluster increases. The nature of the weak interactions in the complexes was characterized by means of atoms in molecules (AIM) and the natural bond orbital (NBO) analyses.  相似文献   

15.
The catalytic properties of the complexes (RCp)2ZrCl2 (R=H, Me, Pri, Bun, Bui, Me3Si,cyclo-C6H11), and Me2SiCp*NBuiZrCl2 (Cp*=C5(CH3)4) combined with the AlBui 3−CPh3B(C6F5)4 cocatalyst in ethylene polymerization were studied. The specific activity of the substituted bis-cyclopentadienyl complexes decreases in the sequence: Me>Pri>Bun>Bui>Me3Si>cyclo-C6H11, which corresponds to the activity sequence for these complexes activated by polymethylaluminoxane (MAO) but is 4–20 times lower in absolute value. Comparison of the polyethylene samples obtained in the presence of the same complexes with MAO and AlBui 3−CPh3B(C6F5)4 cocatalysts showed that polyethylene with much higher molecular mass, melting point, and crystallinity is formed in the presence of the ternary catalytic systems, and this indicates a different nature of the active sites of the catalytic systems. The effective activation energy of polymerization (≈3.6 kcal mol−1), first order with respect to monomer and ≈0.4 order with respect to organoaluminum component, was found for the (PriCo)2ZrCl2−AlBui 3−CPh3B(C6F5)4 catalytic system. It was proposed on the basis of the kinetic data that AliBu3 enters into the composition of the active site to form a bridged heteronuclear cationic complex. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp 301–307, February, 2000.  相似文献   

16.
Hexanuclear oxo titanium(IV) isopropoxide carboxylates, of the general formula [Ti6O6(OPri)6(O2CR)6] (R = But (1), CH2But (2)) and [Ti6O6(OPri)6(O2CC(CH3)2Et)6] · 0.5(C7H8) (3), have been synthesized as polycrystalline powders in order to study their thermal properties and usability as TiO2 CVD precursors. Analysis of thermogravimetric and variable temperature (VT-IR) data shows that the thermal stability of the synthesized complexes decreases as follows 3 > 2 > 1. The composition of the vapors formed during the thermolysis of 13 were qualitatively analysed with VT-IR methods and mass spectrometry (MS-EI). According to obtained results, the decomposition of 1 and 2 proceeds with a partial decomposition and the formation of a volatile and stable titanium species, sufficient for their transport in vapors. The formation of volatile titanium-containing derivatives is an important factor that decides the application of 1 and 2 as precursors in CVD experiments. The high stability of 3 causes the thermal decomposition of this complex to be observed just above 573 K, and volatile titanium-containing derivatives were not detected in vapors. These results indicate that 3 could not be used as a precursor in CVD processes.  相似文献   

17.
Using the ab initio method of SCF MO LCAO
  • 1 SCF MO LCAO: Self-consistent field molecular orbital linear combination of atomic orbitals.
  • in a valency-splitted basis of the Gaussian functions we have studied the addition of various monomers (C3H8, C2H4, C2H2) and dihydrogen to the titanium-alkyl bond in the complex H2TiCH3. The structure of transition states in the insertion reaction, heats of π-complex formation and activation energies for the insertion of the coordinated monomers have been calculated. The calculation results show that the reactivity decreases in the order C2H2 > C2H4 > C3H8 > H2. According to the results obtained, the energy of the π*-antibonding orbital of monomers can serve as an index of relative reactivity in the insertion reaction into the metal-alkyl bond.  相似文献   

    18.
    The ion pair of the stereolabile C3‐symmetric, i+o proton complex [ 1? H]+ of diaza‐macropentacycle 1 and the configurationally stable Δ‐TRISPHAT ([Δ‐ 3 ]?) anion exists in the form of two diastereomers, namely, [Δ‐( 1? H)][Δ‐ 3 ] and [Λ‐( 1? H)][Δ‐ 3 ], the ratio of which, in terms of diastereomeric excess (de) decreases in the order [D8]THF (28 %)>CD2Cl2 (22 %)>CDCl3 (20 %)>[D8]toluene (16 %)>C6D6 (7 %)>[D6]acetone (0 %) at thermodynamic equilibrium. Except in the case of [D6]acetone, the latter is reached after a period of time that increases from 1 h ([D8]THF) to 24 h (CDCl3). Moreover, the initial value of the de of [ 1? H][Δ‐ 3 ] in CDCl3, before the thermodynamic equilibrium is reached, depends on the solvent in which the sample has been previously equilibrated (sample “history”). This property has been used to show that the crystals of [ 1? H][Δ‐ 3 ] formed by slow evaporation of CH2Cl2/CH3OH mixtures had 100 % de, which indicates that [ 1? H][Δ‐ 3 ] has enjoyed a crystallization‐induced asymmetric transformation. Structural studies in solution (NMR spectroscopy) and in the gas phase by calculations at the semiempirical PM6 level of theory suggest that the optically active anion is docked on the i+ (endo) external side of the proton complex such that one of the aromatic rings of [Δ‐ 3 ]? is inserted into a groove of [ 1? H]+, a second aromatic ring being placed astride the outside i+ pocket. Solvent polarity controls the thermodynamics of inversion of the [ 1? H]+ propeller. However, both polarity and basicity control its kinetics. Therefore, the rate‐limiting steps correspond to the ion‐pair separation/recombination and [ 1? H]+/ 1 deprotonation/protonation processes, rather than the inversion of [ 1? H]+, the latter being likely to take place in the deprotonated form ( 1 ).  相似文献   

    19.
    Bending at the valence angle N–Cα–C′ (τ) is a known control feature for attenuating the stability of the rare intramolecular ii hydrogen bonded pseudo five-membered ring C5 structures, the so called 2.05 helices, at Aib. The competitive 310-helical structures still predominate over the C5 structures at Aib for most values of τ. However at Aib, a mimic of Aib where the carbonyl O of Aib is replaced with an imidate N (in 5,6-dihydro-4H-1,3-oxazine = Oxa), in the peptidomimic Piv-Pro-Aib-Oxa (1), the C5i structure is persistent in both crystals and in solution. Here we show that the ii hydrogen bond energy is a more determinant control for the relative stability of the C5 structure and estimate its value to be 18.5 ± 0.7 kJ/mol at Aib in 1, through the computational isodesmic reaction approach, using two independent sets of theoretical isodesmic reactions.  相似文献   

    20.
    The syntheses, physical characterization and crystal structures of two new molecular copper(II) complexes of composition [Cu(C5H5N)2(C7F5O2)2] (1) and [Cu(C5H5N)2(C7F5O2)2(H2O)] (2) (C5H5N = py = pyridine and C7F5O2 = pfb = pentafluorobenzoate) are reported. Single-crystal X-ray structure determinations revealed that in 1, the Cu2+ ion, which lies on a crystallographic inversion centre, is coordinated to two py molecules and two oxygen atoms from two monodentate pfb anions, resulting in a trans-CuN2O2 square planar geometry. In 2, the Cu2+ ion is also coordinated to two py and two pfb species in addition to a water molecule in the apical site of a distorted CuN2O3 square pyramid. In the crystal packing, both 1 and 2 show segregated aromatic π-π stacking interactions in which (py + py) and (pfb + pfb) ring-pairings are seen, but no (py + pfb) pairings occur. Crystal data: 1: C24H10CuF10N2O4, Mr = 643.88, space group , a = 8.0777 (3) Å, b = 8.0937 (3) Å, c = 10.5045 (5) Å, α = 90.916 (3)°, β = 93.189 (2)°, γ = 118.245 (3)°, V = 603.36 (4) Å3, Z = 1. 2: C24H12CuF10N2O5, Mr = 661.90, space group , a = 7.5913 (5) Å, b = 15.6517 (6) Å, c = 21.1820 (14) Å, α = 95.697 (4)°, β = 94.506 (2)°, γ = 91.492 (4)°, V = 2495.2 (3) Å3, Z = 4.  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