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 234 毫秒
1.
The multiphoton ionization of the hydrogen-bonded clusters N,N-dimethylformamide–(methanol)n (DMF–(CH3OH)n) was studied using a time-of-flight mass spectrometer at the wavelengths of 355 and 532 nm. At both wavelengths, a series of protonated DMF–(CH3OH)nH+ ions was obtained. The clusters were also investigated by density functional theory B3LYP method in conjunction with basis sets 6-31+G(d,p) and 6-311+G(2d,p). Equilibrium geometries of both neutral and ionic DMF–CH3OH clusters, and dissociation channels and dissociation energies of the ionic clusters are presented. The results show that when DMF–CH3OH is vertically ionized and dissociated, DMFH+ and CH3O are the dominant products via proton transfer reaction. A high energy barrier makes another channel corresponding to the production of DMFH+ and CH2OH disfavored. In the DMF–(CH3OH)H+ ion, the proton prefers to link with the O atom of DMF molecule. Variation of atomic charges during proton transfer in hydrogen bond of the protonated cluster DMF–(CH3OH)H+ ion is also discussed.  相似文献   

2.
For a closed-shell MO configuration with 2n electrons which occupy n non-degenerate canonical MOs, it is deduced that the RHF energy, Σni=1[2H0nnj-1(2Jij-Kij)], may be expressed in Hückel-like form as 2Σni-1ε, −Σni-1[ji(λ+1)+1,(λ+2)] with λ=2(n-i). The li(λ+1) and Ii(λ+2) are the ionization potentials for the HOMO ψ, which arises after λ and λ+1 electrons have been successively removed from the initial configuration.  相似文献   

3.
The dissociation rates of phenetole ions have been measured as a function of the ion internal energy by the method of photoelectron photoion coincidence spectrometry. The loss of ethylene to produce the phenol ion is the only dissociation pathway from its onset at 9.17 eV up to at least 12 eV. An activation energy of 1.64 ± 0.06 eV with an assumed activation entropy of + 1.8 ± 4 cal/mol-K is derived from fitting the statistical theory decay rates to the measured rates. The transition state energy lies ≈ 1 eV below the thermochemical dissociation limit to the C2H+5 + C6H5O· products. It is thus unlikely that an ion-radical complex is involved in the production of the phenol ion.  相似文献   

4.
The results of theoretical studies on structures and energetics are presented for proton-bound complexes N2H+–XH, N2H+–X2, and N2H+–XY(YX) (X=Y=F, Cl, and Br). In all the monocations complexes, the halogen atom shares a proton with N2. The calculated energetic results show that the stability decreases when descending in the corresponding periodic table column. The possible proton transfer dissociation processes of N2H++XH, N2H++X2, and N2H++XY systems into XH2+, X2H+, XYH+, and YXH+ and molecular N2 are calculated to be endothermic for share of the processes. The NBO results show that the largest intermolecular charge transfer is found in the Br bonded complexes.  相似文献   

5.
Dialkyl disulfide-linked naphthoquinone, (NQ-Cn-S)2, and anthraquinone, (AQ-Cn-S)2, derivatives with different spacer alkyl chains (Cn: n = 2, 6, 12) were synthesized and these quinone derivatives were self-assembled on a gold electrode. The formation of self-assembled monolayers (SAMs) of these derivatives on a gold electrode was confirmed by infrared reflection-absorption spectroscopy (IR-RAS). Electron transfer between the derivatives and the gold electrode was studied by cyclic voltammetry. On the cyclic voltammogram a reversible redox reaction between quinone (Q) and hydroquinone (QH2) was clearly observed under an aqueous condition. The formal potentials for NQ and AQ derivatives were −0.48 and −0.58 V, respectively, that did not depend on the spacer length. The oxidation and reduction peak currents were strongly dependent on the spacer alkyl chain length. The redox behavior of quinone derivatives depended on the pH condition of the buffer solution. The pH dependence was in agreement with a theoretical value of E1/2 (mV) = E′ − 59pH for 2H+/2e process in the pH range 3–11. In the range higher than pH 11, the value was estimated with E1/2 (mV) = E′ − 30pH , which may correspond to H+/2e process. The tunneling barrier coefficients (β) for NQ and AQ SAMs were determined to be 0.12 and 0.73 per methylene group (CH2), respectively. Comparison of the structures and the alkyl chain length of quinones derivatives on these electron transfers on the electrode is made.  相似文献   

6.
Likely candidates for the lowest potential energy minima of (C60)nCa2+, (C60)nF and (C60)nI clusters are located using basin-hopping global optimisation. In each case, the potential energy surface is constructed using the Girifalco form for the C60 intermolecular interaction, an averaged Lennard–Jones C60–ion interaction, and a polarisation potential, which depends on the first few non-vanishing C60 multipole polarisabilities. We find that the ions generally occupy the interstitial sites of a (C60)n cluster, the coordination shell being tetrahedral for Ca2+ and F. The I ion has an octahedral coordination shell in the global minimum for (C60)6I, however for 12  n  8 the preferred coordination geometry is trigonal prismatic.  相似文献   

7.
Poly(methyl methacrylate) (PMMA) modified titanium and zirconium n-butoxide–ethyl acetoacetate (EAA) complex [M5-Ti(OBun)2(EAA)2 and M5-Zr(OBun)2(EAA)2] were obtained from trialkoxysilane-functional PMMA and EAA modified titanium and zirconium alkoxide via the sol–gel method. Infrared (IR), 13C nuclear magnetic resonance (NMR) spectroscopy, and thermogravimetric analysis (TGA) were used to analyze the structures and properties of the hybrids with various proportions of metal oxide species. The effect of the complex of metal oxides and EAA ligands on structure and thermo-oxidative degradation of the M5-Ti(OBun)2(EAA)2 and M5-Zr(OBun)2(EAA)2 hybrids were investigated in this study. The 1H spin–diffusion path length of the hybrids was in a nanometer scale as estimated from the spin–lattice relaxation time in a rotating frame (TH). The apparent activation energies (Ea), evaluated by van Krevelen’s method, for random scission of PMMA segments in hybrids decreased with increasing metal oxide content.  相似文献   

8.
Classical trajectory simulations are performed to study energy transfer in collisions of protonated triglycine (Gly)(3) and pentaglycine (Gly)(5) ions with n-hexyl thiolate self-assembled monolayer (SAM) and diamond [111] surfaces, for a collision energy E(i) in the range of 10-110 eV and a collision angle of 45 degrees. Energy transfer to the peptide ions' internal degrees of freedom is more efficient for collision with the diamond surface; i.e., 20% transfer to peptide vibration/rotation at E(i) = 30 eV. For collision with diamond, the majority of E(i) remains in peptide translation, while the majority of the energy transfer is to surface vibrations for collision with the softer SAM surface. The energy-transfer efficiencies are very similar for (Gly)(3) and (Gly)(5). Constraining various modes of (Gly)(3) shows that the peptide torsional modes absorb approximately 80% of the energy transfer to the peptide's internal modes. The energy-transfer efficiencies depend on E(i). These simulations are compared with recent experiments of peptide SID and simulations of energy transfer in Cr(CO)(6)(+) collisions with the SAM and diamond surfaces.  相似文献   

9.
The enthalpies of formation of the complexes between the silver(I) ion and some sulphur-containing aminopyridines of general formula N(CH2)n−1-S-(CH2)m-NH2 where n = 1, and m = 1,2; 1,3; 2,2; 2,3 have been determined by direct calorimetric titration at 25°C in 0.5 M (K)NO3 solution. The corresponding entropy terms, ΔS, have been calculated using the obtained enthalpy values and the previously reported ΔG values.

In acid medium (pH < 3) coordination occurs through the thioether group and the protonated species AgLH3+2 and AgL2H5+4 are enthalpy-stabilized and entropy-destabilized. At higher pH values (pH &>; 3) additional chelation through the pyridine nitrogen is obvious by a marked increase in the enthalpy of formation of the complexes AgL2H4+3, AgL2H3+2 and AgLH2+. The last complex dimerizes into a cyclic dimer Ag2L, H4+2. At still higher pH values (pH &>; 6) the participation of the amino group in the dimeric chelates Ag2L2H3+ and Ag2L2+2 is revealed again by a marked increase in the heat of complexation.  相似文献   


10.
The density functional theory and Hartree–Fock methods were used to investigate the proton transfer reaction for a series of model clusters of zeolite/(H2O)n; n=1,2,3, and 4. Without promoted water, the hydrogen-bonded dimer of the water/zeolite system exists as a simple hydrogen-bonded complex, ZOH.(H2O)2, and no proton transfer occurs from zeolite to water. The third promoted water, ZOH(H2O)2H2O, was found to induce a pathway for proton transfer, but at least addition two promoted molecules, ZO(H3O+)H2O(H2O)2, must be involved for complete proton transfer from zeolite to H2O. The results show that the hydronium ion in water cluster adsorbed on zeolite, ZO(H3O+)(H2O)3, can considerably affect the structure and bonding of the hydrogen-bonded dimer of water. The OO distance is contracted from 2.818 Å found in the neutral complex, ZOH(H2O)4, to 2.777 Å for ion-pair complex, ZO(H3O+)(H2O)3. The distance between the oxygen of the hydronium ion and the zeolitic acid site oxygen is predicted to be 2.480 Å which is in good agreement with the experimentally observed value of 2.510 Å. The corresponding density functional adsorption energy of the high coverages of adsorbing molecules on zeolite is calculated to be −9.14 kcal/mol per molecule at B3LYP/6-311+G(d,p) level of theory and compares well with the experimental observation of −8.20 kcal/mol.  相似文献   

11.
The electrochemical behaviour of a series of Mo2Cl4(PR3)4 complexes (PR3 = PMe3, PEt3, PPrn3,PBun3, PH2Ph, PMe2Ph, PEt2Ph, PHPh2, PMePh2, PEtPh2, P(OMe)3, P(OMe)Ph2) has been examined by cyclic voltammetry in dichloromethane solution. The phosphines were chosen to provide a wide range of Lewis basicity/π acidity as reflected by Tolman's co IR and Bodner's Δδco 13C NMR spectral parameters for Ni(CO)3(PR3). The Mo2 compounds undergo either quasi-reversible or irreversible one-electron oxidations except for P(OMe)3 and P(OMe)Ph2 for which no clectroactivity was observed before the solvent limit. The anodic peak potentials, Ep,a, span a range of nearly 700 mV. The half-wave potentials, E1/2,for the quasi-reversible couples and Ep,a for all were plotted against the IR and NMR values and against the δ → δ* transition energies for the Mo2 species in dichloromethane and in the solid state. For the organometallic spectral parameters excellent linear correlations were obtained while with the electronic spectral data fair correlations resulted. These results indicate that the Mo2Cl4(PR3)4 complexes become more difficult to oxidize as the electron-withdrawing nature of the PR3 substituents increases and the δ → δ* band energy decreases.  相似文献   

12.
采用高温固相反应,以NaF作助熔剂,在1000 ℃的温度下合成了锕系元素Pu的模拟固化体(Gd1-xCex)2Zr2O7+x (0 ≤ x ≤ 0.7).研究了模拟固化体的物相、热膨胀系数(TEC)、热导率(TC)随温度及组成的变化规律.粉末X射线衍射(XRD)测试结果表明: Gd2Zr2O7基质本身呈弱有序烧绿石结构,而用Ce4+取代Gd3+的模拟固化体都呈缺陷萤石结构. (Gd1-xCex)2Zr2O7+x的Ce(3d) X射线光电子能谱(XPS)有六个峰,结合能分别位于881.7, 888.1, 897.8, 900.4, 907.1, 916.1 eV处,与CeO2的XPS图谱非常相似,说明Ce为四价.随着温度的升高,所有样品的热膨胀系数总体上呈增大趋势.在室温至750 ℃附近,大部分样品的热导率随温度的升高而降低,之后热导率又呈小幅上升.在相同温度下,固化体(Gd1-xCex)2Zr2O7+x (0 ≤ x ≤ 0.7)的热膨胀系数及热导率随组成变化呈相同趋势:在0 ≤ x ≤ 0.1范围内随x的增大而增大,随后在x = 0.1-0.7时逐渐减小.  相似文献   

13.
Closo-BnHn−2(CO)2 (n = 5–12), isolobal analogues of closo-C2Bn−2Hn, have been investigated at the B3LYP/6-311+G**density functional level of theory. The most stable isomers of closo-BnHn−2(CO)2 are similar to those of closo-C2Bn−2Hn in geometric patterns apart from closo-B6H4(CO)2, and closo-BnHn−2(CO)2 is much less strained than closo-C2Bn−2Hn. Energetic analysis identifies closo-B6H4(CO)2, closo-B12H10(CO)2 and closo-B10H8(CO)2 to be most stable, of which the latter two cages have been prepared experimentally. On the basis of the negative and rather large nucleus independent chemical shifts (NICS), closo-BnHn−2(CO)2 are aromatic. To aid further experimental study, the CO stretching frequencies have been computed.  相似文献   

14.
The photophysics of three complexes of the form Ru(bpy)3−(pypm)2+ (where bpy2,2′-bipyridine, pypm 2-(2′-pyridyl)pyrimidine and P=1, 2 or 3) was examined in H2O, propylene carbonate, CH3CN and 4:1 (v/v) C2H5OH---CH3OH; comparison was made with the well-known photophysical behavior of Ru(bpy)32+. The lifetimes of the luminescent metal-to-ligand charge transfer (MLCT) excited states were determined as a function of temperature (between −103 and 90 °C, depending on the solvent), from which were extracted the rate constants for radiative and non-radiative decay and ΔE, the energy gap between the MLCT and metal-centered (MC) excited states. The results indicate that *Ru(bpy)2(pypm)2+ decays via a higher lying MLCT state, whereas *Ru(pypm)32+ and *Ru(pypm)2(bpy)2+ decay predominantly via the MC state.  相似文献   

15.
Reaction of phenyl magnesium bromide with the ,β-unsaturated ketone 3-methyl-2,3,4,5,6,7-hexahydroind-8(9)-en-1-one, followed by an aqueous work-up, generates the pro-chiral tetra-substituted cyclopentadiene, 1-phenyl-3-methyl-4,5,6,7-tetrahydroindene, CpH, a precursor to the η5-cyclopentadienyl ligand in (Cp)2Fe and [(Cp)Fe(CO)]2(μ-CO)2. Both complexes were generated as mixtures of rac-(RR and SS)- and meso-(RS)-isomers, and in either case pure meso-isomer was isolated by crystallisation and characterised by single crystal X-ray structure, both molecules having crystallographic Ci symmetry. Reduction with Na/Hg cleaves meso-(RS)-[(Cp)Fe(CO)]2(μ-CO)2 and the resulting mixture of (R)- and (S)-[(Cp)Fe(CO)2] anions reacts with MeI to give racemic (Cp)Fe(CO)2Me, which was characterised by the X-ray crystal structure. The Cp ligand is more electron donating than (η-C5H5) as revealed by the reduction potential of the (Cp)2Fe+/(Cp)2Fe couple, E°=−0.127 V (vs. Ag  AgCl). Reaction of LiCp with ZrCl4 yields the zirconocene dichloride [Zr(Cp)2Cl2] as mixture of rac- and meso-isomers, from which pure rac-isomer is obtained as a mixture of RR and SS crystals by recrystallisation. The reaction of rac-[Zr(Cp)2Cl2] with LiMe gives rac-[Zr(Cp)2Me2]. The structures of RR-[Zr(Cp)2Cl2] and rac-[Zr(Cp)2Me2] have been determined by X-ray diffraction. The structural studies reveal the influence of the bulky substituted cyclopentadienyl ligand on the metal---Cp distances and other metric parameters.  相似文献   

16.
The Flory–Huggins interaction parameter χ(ri) is considered as dependent on the chain length of a polymer. Therefore, a modified free energy expression of Flory–Huggins theory is obtained for the polydisperse polymer solutions. Based on this modified free energy expression and the thermodynamics of Gibbs, the expression of spinodal for polydisperse polymer solutions is obtained. For a given χ(ri) according to de Gennes, the spinodals are calculated for polydisperse polymer solutions at different molecular weights and their distributions. It is found that all the interested variables rn, rw, rz and molecular weight distribution have an effect on the spinodal for polydisperse polymer solutions, where the effect of changing rw is much greater than that of changing rn, rz and molecular weight distribution.  相似文献   

17.
Ab initio (HF/3-21G*), DFT (B3LYP with basis sets 6-31G*, 6-311+G* and 6-311+G(2d)) and, in some cases, MP2/6-31G* calculations, were done on cyclic dimers, trimers, etc. and on acyclic oligomers (with OH and H on the ends) of sulfur monoxide and sulfur dioxide. The four cyclic (SO)n molecules were (S–O)2 (1,3,2,4-dioxadithietane, 1a), (S–O)3 (1,3,5,2,4,6-trioxatrithiane, 2a), (S(=O))4 (tetrathietane 1,2,3,4-tetraoxide, 1b), and (S(=O))6 (hexathiane 1,2,3,4,5,6-hexaoxide, 2b). The four cyclic (SO2)n molecules were the dioxide of 1a (1,3,2,4-dioxadithietane 2,4-dioxide, 1c), the trioxide of 2a (1,3,5,2,4,6-trioxatrithiane 2,4,6-trioxide, 2c), the tetraoxide of 1b (tetrathietane 1,1,2,2,3,3,4,4-octaoxide, 1d) and the hexaoxide of 2b (hexathiane 1,1,2,2,3,3,4,4,5,5,6,6-dodecaoxide, 2d). The 16 acyclic molecules (oxides of disulfane, trisulfane, etc. and oxides of oxadisulfane, dioxatrisulfane, etc.) were (–S–O–)n, (–S(=O)–)n, (–S(=O)O–)n, and (–S(=O)2–)n, with n from 2 to 5 and HO, H at the ends. Most of these species are relative minima on the B3LYP/6-31G* potential energy surface. In energy content, the SO dimer, etc. lie below, and the SO2 dimer, etc. above, their SOx components, at all the electron-correlated levels.  相似文献   

18.
The reactions of BrMn(CO)5 with the non-chelating stereochemically rigid bidentate ligands (L-L) 1,3-, and 1,4-diisocyanobenzene, 4,4′-diisocyanobiphenyl, and 4,4′-diisocyanodiphenylmethane afford well characterized complexes of the types BrMn(CO)4(L-L), BrMn(CO)3(L-L)2, and [BrMn(CO)4]2(L-L). Similar reactions with [RC5H4Mn(CO)2NO]+PF6 gave mixtures of oligomers of the type [(RC5H4MnNO)n(L-L)n+1]n+[PF6]n.  相似文献   

19.
采用共沉淀法制备了高比表面积的MnxCo3−xO4球形催化剂,研究了NH3选择性催化还原NOx性能。Mn-Co金属氧化物具有尖晶石结构,随着Co含量的增加,晶体结构由四方相转变为立方相。高浓度的表面活性氧物种和变价元素的强有效电子转移(Co3+ + Mn3+ ↔ Co2+ + Mn4+)有利于提高MnxCo3−xO4 (x = 1.0、1.5、2.0)尖晶型石催化剂的氧化还原能力,催化剂表面的Mn富集作用形成了氧缺陷结构和丰富的表面活性位点,进一步促进SCR脱硝反应,呈现出优异的催化性能。Cotet(CoMn)octO4晶体结构中,Mn离子(Mn3+和Mn4+,以三价锰为主)和部分Co离子被配置到八面体中心,这些物种作为活性位点存在着较强的电子转移交互作用,该构型对促进低温脱硝活性和保护活性位点耐受SO2毒害具有重要的意义。Mn-Co尖晶石表面的NH3-SCR脱硝反应过程主要遵循Eley-Rideal反应机理,即吸附态NH3与气态NO (或NO2)的反应路径。随着反应温度的增加,反应生成的NH4NO3中间体很可能转化为NH4NO2物种,进而分解为N2,提高了催化剂的氮气选择性。  相似文献   

20.
The gas-phase stabilities of cluster ions SF+m (SF6)n with m = 0−5 were determined by using a high pressure mass spectrometer. The bond energies of SF+m (SF6)1 were found to be less than 10 kcal/mol and to decrease with m = 0 → 5. There appear to be rather large gaps in the bond energies between n = 1 and 2 for the clusters SF+m (SF6)n with m = 0−4. The structures of SF+5, SF+ (SF6)1, SF+3 (SF6)1, and SF+5 (SF6)1 were investigated by ab initio molecular orbital calculations. For SF+5, the D3h geometry is found to be most stable andC4v is a transition state of the Berry pseudorotation. For the ion-molecule complexes, the “on-top hat” models were found to be the most stable structures.  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