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.
Yield and selectivity of benzene produced from aquasonolysis of the selected cyclic C6Hx hydrocarbons, i.e., 1,4-cyclohexadiene, 1,3-cyclohexadiene, cyclohexene, cyclohexane, and methylcyclopentane, have been investigated in this work. Benzene cannot be detected during the aquasonolysis of cyclohexane and methylcyclopentane. The order of yield and selectivity of benzene was as follows: 1,4-cyclohexadiene>1,3-cyclohexadiene>cyclohexene. The initial concentrations of substrates can affect the yield of benzene. During the aquasonolysis of 1,3-cyclohexadiene and cyclohexene, other C6 species except benzene were also found. It was suggested that benzene could directly be generated by formal dehydrogenation of cyclic C6Hx hydrocarbons.  相似文献   

2.
Aquasonolysis rates and products of selected cyclic C(6)H(x) hydrocarbons, benzene, 1,3-cyclohexadiene, 1,4-cyclohexadiene, cyclohexene, cyclohexane, and methylcyclopentane have been investigated. The sonolysis of selected compounds in aqueous solution follows first-order kinetics, and the aquasonolysis rate correlated well with the water solubility. The degradation rate decreased with the increase of initial concentration. The effect of initial concentration on the degradation of cyclohexene was more significant than that of benzene. The transfer process of organic solutes between cavitation bubbles and the bulk liquid affects the rates and products of their aquasonolysis.  相似文献   

3.
Max Montano 《Surface science》2006,600(9):1809-1816
A scanning tunneling microscope that can be operated in ultra high vacuum (<10−9 Torr) as well as at high pressures (1 − 103 Torr) has been utilized to study the structures formed by cyclic C6 hydrocarbons adsorbed on a platinum (1 1 1) crystal surface. Catalytic reactions of cyclohexene were also studied in the presence of hydrogen at pressures (up to 200 mTorr) and 300 K-350 K temperature range. Cyclohexane and cyclohexene produced the same adsorbed structure, which is attributed to the partially dehydrogenated π-allyl (C6H9). 1,3-Cyclohexadiene produced structures similar to those produced by benzene. In contrast 1,4-cyclohexadiene forms a structure that we attribute to intact molecular 1,4-cyclohexadiene. During reaction the STM images appear disordered, indicative of rapid diffusion of surface species. Addition of 5 mTorr of CO stops the catalytic activity and forms an ordered structure on the surface.  相似文献   

4.
C.D. MacPherson  D.Q. Hu  M. Doan  K.T. Leung   《Surface science》1994,310(1-3):231-242
Recently, we reported a thermal desorption study on the evolution of an intense mass 78 profile for the room-temperature exposure of cyclohexene to Si(111)7 × 7 surface, which was believed to give rise to the formation of benzene by a surface dehydrogenation reaction. Because mass 78 was also found to be the base ion in the gas-phase cracking patterns of both 1,3- and 1,4-cyclohexadiene, the dehydrogenation of cyclohexene on clean, sputtered and oxidized Si(111)7 × 7 surfaces has been re-examined in order to determine the origin of the intense mass 78 desorption profile; i.e. whether it was in fact due to the evolution of benzene or cyclohexadiene, or both. Moreover, a similar dehydrogenation reaction giving rise to toluene desorption between 350 and 600 K has been observed for the room-temperature exposure of 1-methyl-1,4-cyclohexadiene to clean and sputtered Si(111)7 × 7 surfaces. The effects of methyl substitution on the reactivity of these cyclic olefins towards Si(111)7 × 7 can be inferred from these studies. Furthermore, the catalytic activity of Si(111)7 × 7 was found to be enhanced significantly by extending the thermal desorption cycles to a higher temperature of 925 K. The dehydrogenation of these olefins on Si(111)7 × 7 also gave rise to a unique 7 × 1 low energy electron diffraction pattern. Possible factors that may play a role in any proposed model for the dehydrogenation reaction are discussed. Finally, evidence of other surface reactions including cyclohexene hydrogénation to cyclohexane will also be presented.  相似文献   

5.
The adsorption of 1,3-cyclohexadiene, 1,4-cyclohexadiene, cyclohexene and cyclohexane on Pt(1 1 1) was studied using ab initio density functional theory. For 1,3-cyclohexadiene three adsorption modes were distinguished: bridge 1,2-di-σ/3,4-π, hollow 1,4-di-σ/2,3-π and bridge 1,4-di-σ/2,3-π with adsorption energies of −155, −147 and −75 kJ/mol, respectively. Three stable adsorption modes were also identified for 1,4-cyclohexadiene: bridge quadra-σ, hollow di-σ/π and bridge di-π with adsorption energies of −146 kJ/mol, −142 kJ/mol and −88 kJ/mol, respectively. Cyclohexene was found to adsorb in six modes: 4 di-σ and 2 π-adsorption modes. The preferred configuration was found to be boat di-σ with an adsorption energy of −81 kJ/mol. The three other di-σ adsorption modes have comparable adsorption energies, ranging from −64 to −69 kJ/mol. Molecular strain and CPt bonding energies are used to elucidate stability trends. Cyclohexane is found to adsorb only at the hollow site whereby the axial hydrogen atoms are positioned over surface Pt-atoms with an adsorption energy of −37 kJ/mol. The calculations correctly predict the weakening of the axial CH bonds and provide a possible explanation for the large shift in the vibrational frequencies.  相似文献   

6.
The adsorption reactions and binding configurations of cyclohexene, 1,3-cyclohexadiene and 1,4-cyclohexadiene on Si(1 1 1)-7 × 7 were studied using high-resolution electron energy loss spectroscopy (HREELS), ultraviolet photoelectron spectroscopy (UPS), X-ray photoelectron spectroscopy (XPS) and DFT calculation. The covalent attachments of these unsaturated hydrocarbons to Si(1 1 1)-7 × 7 through the formation of Si–C linkages are clearly demonstrated by the observation of the Si–C stretching mode at 450–500 cm−1 in their HREELS spectra. For chemisorbed cyclohexene, the involvement of πC=C in binding is further supported by the absence of C=C stretching modes and the disappearance of the πC=C photoemission. The chemisorption of both 1,3-cyclohexadiene and 1,4-cyclohexadiene leads to the formation of cyclohexene-like intermediates through di-σ bonding. The existence of one πC=C bond in their chemisorbed states is confirmed by the observation of the C=C and (sp2)C---H stretching modes and the UPS and XPS results. DFT calculations show that [4 + 2]-like cycloaddition is thermodynamically preferred for 1,3-cyclohexadiene on Si(1 1 1)-7 × 7, but a [2 + 2]-like reaction mechanism is proposed for the covalent attachment of cyclohexene and 1,4-cyclohexadiene.  相似文献   

7.
Superconductivity in the MgB2 superconductor is described within the framework of a two-band Eliashberg formalism. Different gaps are assumed to open on the different parts of the Fermi surface of this compound. Separation of the order parameter (OP) into two components is achieved by taking the Fourier transform of the OP using the momentum states of the σ- and π-bands of MgB2. Expressions for the Tc and the ratio 2Δσ(0)/kBTc for this superconductor are obtained. Numerical values for these two properties are obtained for a range of values of the cut-off frequency of the phonons responsible for the superconductivity and for a range of values of the ratio between the two energy gaps. This was done for various values of the normalized partial densities of states on the σ-sheet of the Fermi surface.  相似文献   

8.
F. Calaza 《Surface science》2007,601(3):714-722
The adsorption of ethylene on gold-palladium alloys formed on a Pd(1 1 1) surface is investigated using a combination of temperature-programmed desorption (TPD) and reflection absorption infrared spectroscopy (RAIRS). Various alloy compositions are obtained by depositing four monolayers of gold on a clean Pd(1 1 1) surface and annealing to various temperatures. For gold coverages greater than ∼0.7, ethylene adsorbs primarily on gold sites, desorbing with an activation energy of less than 55 kJ/mol. At gold coverages between ∼0.5 and ∼0.7, ethylene is detected on palladium sites in a π-bonded configuration (with a σ-π parameter of ∼0.1) desorbing with an activation energy of between ∼57 and 62 kJ/mol. Further reducing the gold coverage leads to an almost linear increase in the desorption activation energy of ethylene with increasing palladium content until it eventually reaches a value of ∼76 kJ/mol found for ethylene on clean Pd(1 1 1). A corresponding increase in the σ-π parameter is also found as the gold coverage decreases reaching a value of ∼0.8, assigned to di-σ-bonded ethylene as found on clean Pd(1 1 1).  相似文献   

9.
Electron energy-loss measurements have been performed on the vapours of 1,3-cyclohexadiene and 1,4-cyclohexadiene up to 40 eV. The first valence shell excitations which have been calculated by Mulliken are identified in the spectra. Comparison with the Im (?) spectra of polycrystalline samples corroborate these assignements.  相似文献   

10.
V. M. Bermudez   《Surface science》2003,540(2-3):255-264
Cycloaddition reactions between 1,3-butadiene and the C-terminated SiC(1 0 0)-c(2 × 2) surface have been addressed using quantum-chemical methods. The c(2 × 2) structure consists of ---CC--- bridges between underlayer Si atoms which themselves form Si---Si bonds. Of various possible reaction products, the one formed by a [2 + 4] reaction with the ---CC--- bridge (giving a species resembling 1,4-cyclohexadiene) is the lowest in energy. Density functional calculations for the bare c(2 × 2) surface, using a cluster model with mechanical embedding, gave good agreement with structural parameters obtained in previous fully ab initio studies. Similar calculations for the cycloaddition product and for the transition state gave a reaction energy of −50.3 kcal/mol and an activation energy of +6.1 kcal/mol to form a planar ring structure lying normal to the surface. Detailed results for the frequency and infrared polarization behavior of adsorbate vibrational modes have also been obtained.  相似文献   

11.
The rates and products of the sonochemical reactions of benzene, 1,4-cyclohexadiene, 1,3-cyclohexadiene, cyclohexene, and cyclohexane in selected organic solvents have been investigated. The sonochemical reactions of these educts in the investigated organic solvents follow first-order kinetics. Generally, they are sonicated more rapidly in polar than in non-polar solvent; higher volatility of the solute results in faster sonolysis in the organic solvents. However, the sonication of cyclohexane in n-decane and the sonication of benzene in n-propanol are exceptional cases. Since cyclohexane exhibits a much higher lipophilicity and benzene a much higher hydrophilicity than other educts, it might be more difficult to transfer either educt from the bulk liquid into the cavitation bubbles. In tetrachloroethylene, the reactivity of the tested educts with in situ generated chlorine as well as chlorine-containing radical intermediates can accelerate the rate of sonochemical reactions under the employed conditions. In n-propanol and n-decane, the pyrolysis during the collapse of the cavitation bubbles is the only reaction pathway of sonolysis. In tetrachloroethylene, the pyrolysis during the collapse of the cavitation bubbles and the free radical reaction in the bulk liquid may occur simultaneously. Except for the products generated from sonolysis, products formed from chlorine transformations (substitution or addition reactions) are detected. Benzene is hardly decomposed in tetrachloroethylene. However, when FeCl3 is added into the reaction system, benzene is sonoconverted rapidly, and the product chlorobenzene was detected. In organic solvents, the sonoreaction rates and the sonoproducts are dependent on the physicochemical properties of the solvents used, as well as the volatility, the polarity and the reactivity of educts.  相似文献   

12.
Fourier transform infrared spectroscopy has been applied to the study of cyclohexane adsorbed on Al2O3 and Pt/Al2O3 surfaces. Earlier studies of benzene on these same materials have also been extended to include benzene adsorbed on a Pt/Al2O3 surface which contains structured carbon residues. The data provide indirect evidence for the formation of a carbon residue on Pt/Al2O3 which retains the six-membered cyclic structure of the parent adsorbates. The carbon residue can be formed upon vacuum heating of the parent C6 ring molecules chemiorbed on Pt/Al2O3. There is spectroscopic evidence that cyclohexane dehydrogenates on Pt/Al2O3 at 300 K to form two different chemisorbed species; a π-bonded benzene and a dissociated σ-bonded benzene. These two chemisorbed species have CH stretching vibrations centered at 3030 and 2947 cm?1, respectively. Benzene added to a clean catalyst surface forms only a π-bonded benzene. However, benzene added to Pt/Al2O3 with ordered carbon residues forms both π- and σ-bonded benzenes. The addition of H2 at 300 K to any of the π- or σ-bonded benzenes or to the carbon residue results in the formation of cyclohexane physisorbed on the catalyst. The absence of CH3 groups upon hydrogenation suggests the lack of CC bond breaking during adsorption or hydrogenation. Simultaneous infrared and thermal desorption studies on chemisorbed deuterated benzene (from C6D12) indicate that the a-bonded species exchange H from the surface OH groups of the alumina support more readily than does the π-bonded benzene. In addition to hydrogen exchange with the support, thermal desorption experiments indicate the oxidation of a portion of the chemisorbed hydrocarbons and/or carbon residue by oxygen from the alumina support. Therefore, the support is capable of playing a direct role in reactions occurring on the catalyst surface.  相似文献   

13.
《Solid State Communications》2002,121(9-10):479-484
The optical conductivity of MgB2 has been determined on a dense polycrystalline sample in the spectral range 6 meV–4.6 eV using a combination of ellipsometric and normal incidence reflectivity measurements. σ1(ω) features a narrow Drude peak with anomalously small plasma frequency (1.4 eV) and a very broad ‘dome’ structure, which comprises the bulk of the low-energy spectral weight. This fact can be reconciled with the results of band structure calculations by assuming that charge carriers from the 2D σ-bands and the 3D π-bands have principally different impurity scattering rates and negligible interband scattering. This also explains a surprisingly small correlation between the defect concentration and Tc, expected for a two-gap superconductor. The large 3D carrier scattering rate suggests their proximity to the localization limit.  相似文献   

14.
Feng Gao 《Surface science》2006,600(9):1837-1848
The chemistry of ethylene adsorbed on a thin MoAl layer grown in ultrahigh vacuum on a thin alumina film is studied using a combination of temperature-programmed desorption and X-ray, Auger and reflection absorption infrared spectroscopies. Both di-σ-bonded and a small proportion of π-bonded ethylene are found, where the di-σ-bonded ethylene has a σ/π parameter of ∼0.8 and a heat of adsorption of ∼70 kJ/mol. The ethylene self-hydrogenates to yield ethane and a small amount of methane is detected. The surface hydrogenation activation energy of di-σ-bonded ethylene is ∼65 kJ/mol, while the π-bonded species hydrogenates more easily. Adsorbed ethyl species grafted onto the surface by decomposing ethyl iodide predominantly undergo β-hydride elimination to yield ethylene. Ethyl species hydrogenate to ethane at a lower temperature than does di-σ-bonded ethylene implying that addition of hydrogen to adsorbed ethylene is slower than the rate of ethyl hydrogenation.  相似文献   

15.
C2H4在Ru(1010)表面吸附与分解的研究   总被引:2,自引:0,他引:2       下载免费PDF全文
用X射线电子能谱(XPS)、热脱附谱(TDS)和紫外光电子能谱(UPS)方法研究了乙烯(C2H4)在Ru(1010)表面的吸附,在低温下(200K以下)乙稀(C24)可以在Ru(1010)表面上以分子状态稳定吸附,在200K以上乙烯(C2H 4)则发生了脱氢分解反应.TDS结果表明乙烯(C2H4)分 解后的主要产物为乙炔(C< 关键词: 乙烯 钌(1010)表面 吸附与分解  相似文献   

16.
Cis and trans proton coupling constants in ethylene have been calculated using the variational procedure and valence bond formalism. The problem is treated as an eight-electron eight-orbital one, with six orbitals belonging to the σ framework and two to the π. σ-π exchange is implicitly included.  相似文献   

17.
Isomer shift (IS) and electric quadrupole splitting (QS) of the 77 keV γ-rays of Au197 were investigated for a large number of Au (I) and Au (III) compounds at 4.2 °K by nuclear γ-resonance Spectroscopy. A close correlation between the observed isomer shifts and the spectrochemical series of the ligands was observed. For each oxidation state, isomer shift and electric quadrupole splitting show approximately a linear relationship. On the basis of LCAO-MO theory, the experimental results are interpreted by covalency effects in the molecular orbitals, synergic coupling of σ- and π-bonds, and the empirically known donor and acceptor properties of the ligands.  相似文献   

18.
The empirical correlation of the photoelectron spectra of 1,4-cyclohexadiene (molecule 4), 1, 4, 5, 8-tetrahydronaphthalene (molecule 5), 1, 4, 5, 6, 9, 1 0-hexahydroanthracene (molecule 6), and 1, 4, 5, 6, 7, 10, 11, 12-octahydronaphthacene (molecules 6) proves that the electronic ground state of these molecules is 2B1u, assuming that they have D2h symmetry. In particular this confirms previous predictions for 1,4-cyclohexadiene (molecule 4), for which the “inverted” orbital sequence 2b1u(π) above lb3g(π) had been proposed under the assumption that hyperconjugative “through-bond” interaction dominates the “through-space” interaction of the two semi-localized π-orbitals.  相似文献   

19.
The adsorption of tetrahydrofuran (THF) and furan (F) on polycrystalline Pt-foil has been studied by work function change (Δφ) measurements and Auger electron spectroscopy (AES). The results were compared with those obtained earlier for the adsorption of cyclic hydrocarbons. The general character of the adsorption process (i.e. the dependence of ∣Δφ∣ and C/Pt ratio on temperature and exposure, and the ∣Δφ∣ vs. C/Pt correlation) is similar to the one observed during adsorption of cyclic hydrocarbons. THF produces σ-bonded adspecies in a similar manner to cyclopentane. F forms first a π-bonded surface complex, as do benzene and toluene. The π-adsorbed species are then transformed into σ-bonded ones under exposure to excess F. The numeric values characterizing the adsorption of the two heterocycles (mainly the C/Pt values) differ substantially from those measured for the adsorption of cyclic hydrocarbons.  相似文献   

20.
Reactions of acrolein, water, and oxygen with the vacuum-reduced surface of TiO2(1 1 0) are reported in a temperature programmed reaction study of the interaction of an aldehydic pollutant with a reducible metal oxide. A total of 25% of the acrolein that binds to the surface is converted to products. Notably, carbon-carbon coupling occurs with 86% selectivity for formation of C6 products: C6H8, identified as 1,3-cyclohexadiene, in a peak at 500 K and benzene immediately thereafter at 530 K. Acrolein is evolved from the surface in three peaks: a peak independent of coverage at 495 K, attributed to decomposition of an intermediate that is partly converted to C6H8; a coverage-dependent peak that shifts from 370 K (low coverage) to 260 K (high coverage), which is attributed to adsorption at 5-fold coordinated Ti sites; and a multilayer state at 160 K. Water and acrolein compete for 5-fold coordinated titanium sites when dosed sequentially. The addition of water also opens a new reaction pathway, leading to the hydrogenation of acrolein to form propanal. Water has no effect on the yield of 1,3-cyclohexadiene. Exposure of the surface to oxygen prior to acrolein dosing quenches the evolution of acrolein at 495 K and concurrently eliminates the coupling. From these results, we propose that reduced subsurface defects such as titanium ion interstitials play a role in the reactions observed here. The notion that subsurface defects may contribute to the reactivity of organic molecules over reducible oxide substrates may prove to be general.  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