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.
The structure and texture characteristics of the hybrid organic–inorganic adsorbents, which were obtained by using of two-component systems of “structure-forming agent/trifunctional silane”, are compared as follows: the first component is Si(OC2H5)4 or (C2H5O)3Si–A–Si(OC2H5)3, where A = –(CH2)2– or –C6H4–; the second one is alkoxysilane with amine (–NH2, NH, –NH(CH2)2NH2) and thiol (–SH) groups. The adsorbents, derived from TEOS, have more accessible functional groups (2.6–4.2 mmol/g) than xerogels, which are based on bis(triethoxysilanes) (1.0–2.6 mmol/g). On another hand xerogels derived from bis(triethoxysilanes) have a more extended porous structure (Ssp =516–968 m2/g, Vs = 0.418–1.490 cm3/g, d = 2.5–15.0 nm) than those that are based on TEOS (Ssp = 4–631 m2/g, Vs = 0.005–1.382 cm3/g, d = 2.3–17.7 nm). The geometric dimensions of functional groups have a more essential effect on the parameters of porous structure in the case of TEOS-derived xerogels. Using solid-state NMR spectroscopy, it has been shown that in synthesis of xerogels with the use of TEOS, the molecular frame of globules is formed by structural units Qn (n = 2,3,4), and the functional groups exist as structural units of Tn (n = 2,3). The xerogels obtained with using bis(triethoxysilanes) consist only of structural units of Tn-type (n = 1,2,3).  相似文献   

2.
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   

3.
The directed oligomerization of propene and 1-hexene was carried out with a series of Cp′(C5H5)ZrCl2 and Cp2′ZrCl2 pre-catalysts (Cp′=C5HMe4, C4Me4P, C5Me5, C5H4tBu, C5H3-1,3-tBu2, C5H2-1,2,4-tBu3) together with (C5H5)2ZrCl2. Oligomers in the molar mass range 300–1500 g/mol for propene and 200–3000 g/mol for 1-hexene were synthesized at 50 °C. The majority of oligomer molecules contain a double-bond end group. Oligomer characterization was carried out by gel permeation chromatography (GPC), 1H and 13C NMR. Vinylidene double bonds (from β-hydrogen elimination) are solely found for the tert-butyl-substituted zirconocenes and for most of the unsymmetrical methyl-substituted Cp′(C5H5)ZrCl2 systems (except Cp′=phospholyl). With (C4Me4P)(C5H5)ZrCl2 and with the symmetrical methyl-containing Cp2′ZrCl2 pre-catalysts, also vinyl end groups (from β-methyl elimination) are observed in the case of oligopropenes. The vinylidene/vinyl ratio depends on the ligand and the vinyl content increases from C5HMe4 (65/35) over C4Me4P (61/39) to C5Me5 (9/91). The phospholyl zirconocenes and (C5HMe4)2ZrCl2 also exhibit chain-transfer to aluminum thereby giving saturated oligomers.  相似文献   

4.
The complexes [Zn2(S2CTR)4] (T = 2,5-disubstituted thiophene, R = C4H9 (1), C6H13 (2), C8H17 (3), C12H25 (4) and C16H33 (5)) have been synthesized and their structural features investigated. Compared to the analogous dithiobenzoate complexes, the crystal structure determination of 2 revealed that the thiophene induces a “step-rod” chain pattern instead of the linear, rodlike structure found for the corresponding dithiobenzoates. Complexes 1–5 did not display mesophases under thermal conditions, but an irregular melting pattern was observed for 3 and 4.  相似文献   

5.
[1,8-C10H6(NR)2]TiCl2 (3; R=SiMe3, SiiBuMe2, SiiPr3) complexes have been prepared from dilithio salts [1,8-C10H6(NR)2]Li2 (2) and TiCl4 in diethyl ether in moderate yields (60–63%). These complexes showed significant catalytic activities for ethylene polymerization and for ethylene/1-hexene copolymerization in the presence of methylaluminoxane (MAO), methyl isobutyl aluminoxane (MMAO), AliBu3– or AlEt3–Ph3CB(C6F5)4 as a cocatalyst. The catalytic activities performed in heptane (cocatalyst MMAO) were higher than those carried out in toluene (cocatalyst MAO): 709 kg-PE/mol-Ti·h could be attained for ethylene polymerization by using [1,8-C10H6(NSiiBuMe2)2]TiCl2–MMAO catalyst system.  相似文献   

6.
Spatial structure of six β-substituted enones, with common structure R1O–CR2CH–COCF3, were R1 = C2H5, R2 = H (ETBO); R1 = R2 = CH3 (TMPO); R1 = C2H5, R2 = C6H5 (ETPO); R1 = C2H5, R2 = 4- O2NC6H4 (ETNO); R1 = C2H5, R2 = C(CH3)3 (ETDO) were investigated by 1H and 19F NMR, infrared spectroscopy and AM1 calculations. NMR spectra revealed that enones (MBO), (ETBO) and (TMPO) are exclusively (3E) isomers, whereas in (ETPO), (ETNO) and especially in (ETDO) the percentage of (3Z) isomers is significant and depends on the nature of solvents. Conformational behaviour of studied enones are determined by the rotation around of CC double bond, C–C and C–O single bonds (correspondingly trifluoroacetyl and alkoxy groups), and (EZZ) conformer being the most stable in all cases. IR spectra revealed that with the exception of (ETDO) (EZZ) conformer is most populated in all cases. Bulky substituents like phenyl or tert-butyl group at β-position of enone result in the equilibrium mainly between (EZZ) and (ZZZ) forms, whereas β-hydrogen and β-methyl substituents determine the equilibrium between (EZZ) and (EEZ) or (EZE) conformers.  相似文献   

7.
From the reaction of MeReO3 with the neutral arylamine C6H5CH2NMe2 and the aryldiamine C6H4(CH2NMe2)2−1,3, have been isolated in good yields the 1/1 adduct complex [MeReO3 · C6H5CH2NMe2], 1, and the 2/1 adduct complex [(MeReO3)2 · C6H4(CH2NMe2)2− 1,3], 2, respectively. The X-ray molecular structure of 2 shows that both rhenium centres have a trigonal bipyramidal geometry and in the axial positions of each rhenium centre are one of the NMe2 units of the aryldiamine ligand and a methyl group. The mono(ortho)-chelated arylaminorhenium trioxide complex [ReO3(C6H4CH2NMe2−2], 3, can be synthesized by a transmetallation reaction of ClReO3 with [ZnC6H4CH2NMe2−22] in a 2:1 molar ratio. In a similar way the bis(ortho)-chelated arylaminorhenium trioxide complex [ReO3C6H3(CH2NMe2)2−2,6], 4, can be synthesized by addition of a mixture of [Li2C6H3(CH2NMe2)2−2,62] and ZnCl2 to ClReO3. Complexes 3 and 4 have been isolated as white solids in 66% and 81% yields respectively. The rhenium centre in complex 4 has a bicapped tetrahedral geometry in which the monoanionic C6H3(CH2NMe2)2−2,6 ligand is pseudo-facially bonded with a characteristic N1-Re-N2 angle of 107.7(3)°, a Re-Cipso bond length of 2.112(11) Å and Re-N1 and Re-N2 bond lengths of 2.518(9) Å and 2.480(8) Å respectively.  相似文献   

8.
The pyrolysis mechanism of important intermediate 1-hexene of carbon matrix precursor cyclohexane was studied theoretically. Possible reaction paths were designed based on the potential surface scan and electron structure of the initial C–C bond breaking reactions. Thermodynamic and kinetic parameters of the possible reaction paths were computed by UB3LYP/6-31+G* at different temperature ranges. The results show that 1-hexene pyrolyzes at 873 K. When below 1273 K, the major reaction paths are those that produce C3H4, and above 1273 K, the major reaction paths are those that produce C3H3 from the viewpoint of thermodynamics. From the viewpoint of kinetics, the major product is C3H3, it results from the pyrolysis reaction of 1-hexene cracking bond C3–C4 and generating C3H5 and C3H7 with the activation energy ΔE0θ=296.32 kJ/mol. Kinetic results also show that product C3H4 accompany simultaneously, which is the side reaction starting from the pyrolysis of 1-hexene forming C4H7 and C2H5 with the activation energy of 356.73 kJ/mol. When reaching 1473 K, the rate constant of the rate-determining steps of these two reaction paths do not show much difference, which means both the reaction paths exist in the pyrolysis process at the high temperature. The above results are basically in accordance with mass spectrum analysis and far more specific.  相似文献   

9.
Reaction of R---N=C=N---R (R=p-Me-C6H4) and R---P==C=P---R (R=2,4,6-tBu3C6H2) with the di-iron aminocarbene complex [Fe2(CO)7{1μ-C(Ph)C(NEt2)}] (1c) gave corresponding complexes [Fe2(CO)6{C(Ph)C(NEt2)C(NC6H4Me)N (C6H4Me)}] (2) and [Fe2(CO)6{C(Ph)C(NEt2)C(PC6H2tBu3)P(C6H2tBu3)}] (4), resulting from a coupling reaction with carbon-carbon bond formation. [Fe2(CO)5(CNC6H4Me){C(Ph)C(NEt2)N(C6H4Me)}], complex 3, obtained in the reaction with R---N=C=N---R, resulted from C=N bond rupture insertion of a nitrene fragment into the Fe=C bond. Complexes 2–4 were characterized by X-ray diffraction. The different geornetries of complexes 2 and 4 are discussed. The formation of these complexes may be explained by cycloaddition on the Fe =C metal-carbene bond.  相似文献   

10.
Two organogold derivatives of diphenylmethane and diphenylethane, Ph3PAu(o-C6H4)CH2(C6H4-o)AuPPh3 (1) and Ph3PAu(o-C6H4)(CH2)2(C6H4-o)AuPPh3 (2), have been synthesized by the reaction of ClAuPPh3 with Li(o-C6H4)CH2(C6H4-o)Li and Li(o-C6H4)(CH2)2(C6H4-o)Li respectively. The interaction of 1 with dppe results in the replacement of the two PPh3 groups to give a macrocyclic compound (3) that includes an Au Au bond. Compounds 1 and 2 react with one or two equivalents of [Ph3PAu]BF4 to form new types of cationic complex [CH2(C6H4-o)2(AuPPh3)3]BF4 (4), [CH2(C6H4-o)2(AuPPh3)4](BF4)2 (5), and [(CH2)2(C6H4-o)2(AuPPh3)4](BF4)2 (6). Complexes 1–6 have been characterized by X-ray diffraction studies, FAB MS, and IR as well as by 1H and 31P NMR spectroscopy. A complicated system of Au H-C agostic interactions, involving the bridging alkyl groups (—CH2— and CH2-CH2—) of diphenylmethane and diphenylethane ligands, has been found to occur in complexes 1–3 and 6.  相似文献   

11.
We investigated catalytic behavior of iron in CO2 hydrogenation with and without a ruthenium component. Calcined iron-based catalysts were reduced by H2 and characterized by XRD, BET surface area and CO2, CO and C2H4 temperature-programmed desorption (TPD), and tested for CO2 hydrogenation. When Fe-K/γ-Al2O3 was used as a catalyst, CO2 conversion was 36%, but when Fe-Ru-K/γ-Al2O3 was used, CO2 conversion was 41%. The product selectivities for catalysts with and without the ruthenium component were also compared. Fe-K/γ-Al2O3 exhibited higher methane (16 mol%) and C2–C4 selectivity (39.6 mol%) than Fe-Ru-K/γ-Al2O3. The main products obtained with Fe-Ru-K/γ-Al2O3 were higher hydrocarbons such as C5+ hydrocarbons. For Fe-Ru-K/γ-Al2O3, the product distribution followed the Anderson–Schultz–Flory (ASF) distribution. However, in the case of Fe-Ru-K/γ-Al2O3, the hydrocarbon distribution deviates from the ideal ASF distribution. It is concluded that the readsorption rates of the primary hydrocarbon product increase exponentially with chain length in the ruthenium promoted catalytic system. The behavior of catalysts with and without the ruthenium will be explained by the CO2-, CO- and C2H4– profiles. In this study, it was confirmed that ruthenium component promoted the readsorption ability of -olefin, and then the chain length of hydrocarbon is higher. In addition, the microcrystalline wax produced in CO2 hydrogenation was a high-crystalline and olefin-rich hydrocarbon.  相似文献   

12.
The molecular structures and electron affinities of the C6HCl5 and C6Cl6 molecules have been determined using seven pure Density Functional Theory (DFT) or hybrid Hartree–Fock/DFT methods. The EAs of ten kinds of monochlorobenzene, dichlorobenzene, trichlorobenzene and tetrachlorobenzene are also predicted. The basis set used in this work is of double-ζ plus polarization quality with additional diffuse s- and p-type functions, denoted DZP++. These methods have been carefully calibrated (Chem. Rev. 2002, 102, 231). The geometries are fully optimized with each DFT method independently. The equilibrium configuration of hexachlorobenzene is found to be planar with D6h symmetry. The pentachlorobenzene is planar with C symmetry. Three different types of the neutral-anion energy separations reported in this work are the adiabatic Electron Affinity (EAad), the vertical Electron Affinity (EAvert), and the Vertical Detachment Energy (VDE). The most reliable adiabatic electron affinities of the chlorinated benzenes obtained at the BHLYP level of theory are −0.18 eV (C6H5Cl), 0.07 eV (1,2-C6H4Cl2), 0.07 eV (1,3-C6H4Cl2), 0.04 eV (1,4-C6H4Cl2), 0.29 eV (1,2,3-C6H3Cl3), 0.31 eV (1,2, 4-C6H3Cl3), 0.31 eV (1,3,5-C6H3Cl3), 0.51 eV (1,2,3,4-C6H2Cl4), 0.48 eV (1,2,4,5-C6H2Cl4), 0.50 eV (1,2,3,5-C6H2Cl4), 0.74 eV (C6HCl5) and 0.79 eV (C6Cl6), respectively.  相似文献   

13.
Treatment of the diaminobenzene [C6H4{CH2NMe2}2-1,3] (NCN-H, 1) with one or two equivalents of cis-PtCl2(DMSO)2 leads to exclusive formation of the doubly cycloplatinated species [C6H4{CH2NMe2}2-1,5-{PtCl(DMSO)}2-2,4] (3), which upon addition of triphenylphosphine yields the bisphosphine adduct [C6H4{CH2NMe2}2-1,5-{PtCl(PPh3)}2-2,4] (4). The X-ray molecular structure of 4 revealed the presence of highly distorted square planar Pt(II) centers which is caused by close proximity of the two phosphine donor ligands. Complexes of type 3 can be regarded as suitable starting materials for the directional build-up of larger macromolecular structures.  相似文献   

14.
Characteristics of methyl methacrylate (MMA) polymerization using oscillating zirconocene catalysts, (2-Ph-Ind)2ZrX2 (X = Cl, 1; X = Me, 2), mixtures of rac- and meso-zirconocene diastereomers, (SBI)ZrMe2 [3, SBI = Me2Si(Ind)2] and (EBI)ZrMe2 [4, EBI = C2H4(Ind)2], as well as diastereospecific metallocene pairs, rac-4/Cp2ZrMe2 (5) and rac-4/CGCTiMe2 [6, CGC = Me2Si(Me4C5)(t-BuN)], are reported. MMA polymerization using the chloride catalyst precursor 1 activated with a large excess of the modified methyl aluminoxane is sluggish, uncontrolled, and produces atactic PMMA. On the other hand, the polymerization by a 2/1 ratio of 2/B(C6F5)3 or 2/Ph3CB(C6F5)4 is controlled and produces syndiotactic PMMA. Mixtures of diastereomeric ansa-zirconocenes 3 or 4 containing various rac/meso ratios, when activated with B(C6F5)3, yield bimodal PMMA; this behavior is attributed to the meso-diastereomer that, in its pure form, affords bimodal, syndio-rich atactic PMMA. For MMA polymerization using diastereospecific metallocene pairs, rac-4/5 and rac-4/6, the isospecific catalyst site dominates the polymerization events under the conditions employed in this study, and the aspecific and syndiospecific sites are largely nonproductive, thereby forming only highly isotactic PMMA.  相似文献   

15.
《Polyhedron》1988,7(24):2601-2603
Distibines of the type R2SbSbR′2 with R = CH3, R′ = C2H5 (1), R = CH3, R′= n-C3H7 (2), R = CH3, R′= C6H5 (3), R = C2H5, R′= C6H5 (4), R = n-C3H7, R′ = C6H5 (5), and R = CH3, R′ = 2,4,6-(CH3)2C6H2 (6) are formed in equilibria by exchange reactions of the respective distibines of the type R4Sb2 and R′4Sb2.  相似文献   

16.
The synthesis, crystal structure and magnetic measurements of three new polynuclear tetracarboxylato-bridged copper(II) complexes, i.e. {[Cu4(phen)2(μ-O2CC2H5)8] · (H2O)}n (1), [Cu2(μ-O2CC6H4OH)4(C7H7NO)2] · 6H2O (2) and [Cu2(μ-O2CCH3)4(C7H7NO)2] (3) (phen = 1,10-phenanthroline, O2CC6H4OH = 3-hydroxy benzoate, C7H7NO = 4-acetylpyridine) are reported. All compounds consist of dinuclear units, in which two Cu(II) ions are bridged by four syn,syn11:μ carboxylates, showing a paddle-wheel cage type with a square-pyramidal geometry, arranged in different ways. The structure of compound 1 consists of an one-dimensional structure generated by an alternating classical dinuclear paddle-wheel unit and an unusual dinuclear Cu2(μ-OCOC2H5)2(μ-O2CC2H5)2(phen)2unit, which are connected to each other via a syn,anti-triatomic propionato bridge in an axial-equatorial configuration. The adjacent chains are connected to generate a 2D structure through the face-to-face π–π interaction between phen rings. Structures of compounds 2 and 3 both consist of a symmetric dinuclear Cu(II) carboxylate paddle-wheel core and pyridyl nitrogen atoms of 4-acetylpyridine ligand at the apical position, and just differ in the substituents of the equatorial ligands.

The magnetic properties have been measured and correlated with the molecular structures. It is found that in the two classical paddle-wheel compounds the Cu(II) ions are strongly antiferromagnetically coupled with J = −278.5 and −287.0 cm−1 for complexes 2 and 3, respectively. In compound 1 the magnetic susceptibility could be fitted with two different, independent Cu(II) units, one strongly coupled and one weakly coupled; the paddle-wheel dinuclear unit has the strongest antiferromagetic coupling with a value for J of −299.5 cm−1, whereas the Cu(II) ions in the propionato-bridged dinuclear unit of 1 display a very weak antiferromagnetic coupling with a value for J = −0.75 cm−1, due to the orthogonality of the magnetic orbitals. Also the exchange within the chain is therefore very weak. The magneto-structural correlations for complexes 1, 2, and 3 are discussed on the basis of the structural parameters and magnetic data for the complexes.  相似文献   


17.
Thermal decomposition of mixed ligand thymine (2,4-dihydroxy-5-methylpyrimidine) complexes of divalent Ni(II) with aspartate, glutamate and ADA (N-2-acetamido)iminodiacetate dianions was monitored by TG, DTG and DTA analysis in static atmosphere of air. The decomposition course and steps of complexes [Ni(C5H6N2O2)(C4H5NO4)2−(H2O)2]·H2O, [Ni(C5H6N2O2)(C5H7NO4)2−(H2O)2]·H2O and [Ni(C5H6N2O2)(C6H8N2O5)2−(H2O)2]·1.5H2O were analyzed. The final decomposition products are found to be the corresponding metal oxides. The kinetic parameters namely, activation energy (E*), enthalpy (ΔH*), entropy (ΔS*) and free energy change of decomposition (ΔG*) are calculated from the TG curves using Coats–Redfern and Horowitz–Metzger equations. The stability order found for these complexes follows the trend aspartate > ADA > glutamate.  相似文献   

18.
The oxidative dehydrogenation of ethane over NiO-loaded MgO with high surface area was carried out using a fixed-bed flow reactor at 600 °C under atmospheric pressure.

At 600 °C, the oxidative dehydrogenation of ethane (C2H6/O2 = 1) without dilution with an inert gas resulted in C2H6 conversion of 68.8% and a high C2H4 selectivity of 52.8%, which corresponds to a C2H4 yield of 36.3%. In addition, the catalytic activity did not decrease for at least 10 h. X-ray photoelectron spectra of the catalysts after the reaction exhibited that the initial valence state of Ni2+ (NiO) was maintained during the oxidative dehydrogenation of ethane. However, when NiO-loaded MgO was reduced with H2 prior to the reaction, C2H4 selectivity decreased to nearly zero and high CO and H2 selectivities were observed with the C2H6 conversion of 50 %, indicating that partial oxidation of C2H6 proceeded. Therefore, it seems important to keep Ni species as an oxide phase on the support, and for this purpose, use of the high surface area of MgO is essential.  相似文献   


19.
The encounter complex C2H4…ClF was isolated by using a fast-mixing nozzle before chemical reaction could occur between the components and was characterised by Fourier-transform microwave spectroscopy. Rotational constants, centrifugal distortion constants, Cl nuclear quadrupole constants and Cl spin-rotation constants were determined for the isotopomers C2H435ClF and C2H437ClF. The complex has C2v symmetry with the ClF subunit perpendicular to the plane of C2H4 and oriented so that Cl is closer to C2H4. Both the centrifugal distortion constant ΔJ and the Cl nuclear quadrupole coupling constants indicate that the complex is relatively weakly bound and it is concluded that the interaction between the subunits is largely electrostatic in origin.  相似文献   

20.
Theoretical calculations (DFT, MP2) are reported for up to four sets of reaction products of trimethylphosphine, (CH3)3P, each with H2O, HCl and HF together with DFT calculations on up to three sets of reaction products of substituted phosphonium cations, (CH3)3P–R+. These products comprise (a) P(III) normal complexes (CH3)3PHY, (b) P(IV) ‘reverse’ complexes Y(H–CH2)3P–R, (c) P(IV) ylidic complexes YHCH2(CH3)2P–R and (d) P(V) covalent compounds Y–P(CH3)3–R for Y=HO, Cl and F and R=H, CH3, C2H5, C2H4OH and C2H4OC:OCH3. Calculations are carried out at the B3LYP/6-31+G(d,p) level in all cases and also at the MP2/6-31+G(d,p) level for systems in which R=H. Minimum energy structures are determined for predicted complexes or structures and geometrical properties, harmonic vibrations and BSSE corrected binding energies are reported and compared with the limited experimental information available. Potential energy scans predict equilibria between covalent trigonal bipyramidal P(V) forms and reverse complexes comprising hydrogen bonded or ion pair, tetrahedral P(IV) forms separated by low potential energy barriers. Similar scans are also reported for equilibria between reverse complexes and ylidic complexes for Y=OH and R=CH3, C2H5, C2H4OH and C2H4OC:OCH3. Corrected binding energies, structures and values of harmonic modes are discussed in relation to bonding The names ‘pholine’ and ‘acetylpholine’ are suggested for phosphorus analogues to choline and acetylcholine.  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