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 46 毫秒
1.
The performance of heteronuclear clusters [AlXO3]+ (X=Al, AlO4, AlMg2O2, AlZnO, AlAu2, Mg, Y, VO, NbO, TaO) in activating methane has been explored by a combination of high–level quantum calculations with reported and supplementary gas-phase experiments. With different dopants in [AlXO3]+, the mechanism, reactivity and selectivity towards methane activation varies accordingly. The classic HAT competes with PCET, depending on the composition of intramolecular interactions. Although the existence of terminal oxygen radical is beneficial for classic HAT, the Alt−C interaction in the [AlXO3]+ clusters as enhanced by the strongly electronegative doping groups (X=Al, AlZnO, Mg, Zn, VO, NbO, TaO) favors the PCET process, facilitating C−H bond breaking. In addition, with different dopants, the destiny of the split methyl group varies accordingly. While strong interaction between Alt and CH3 results in the formation of the Alt−C bond, dopants with variable valance may promote the formation of deep-oxidation products like formaldehyde. It has been discussed in detail how to regulate the activity and selectivity of the active center of the catalyst via rational doping.  相似文献   

2.
The potential energy surfaces (PES), energies E, and activation barriers h of elementary reactions of dissociative addition of CH4 and C2H6 molecules to the Al12Ti cluster with a marquee structure in the singlet and triplet states were calculated within the B3LYP approximation of the density functional theory using the 6-31G* basis set. The first stage of the reaction Al12Ti + CH4 leads to the adsorption complex CH4 · Al12Ti with the R(TiC) distance of ~2.4 Å. The methane molecule is coordinated as a tridentate ligand the singlet state and as a bidentate ligand in the triplet state, although both coordination modes are close in energy. In the transition state, the CH4 molecule is coordinated through its active C-H bond to an inclined Ti-Al edge of the cluster, and the C-H bond is significantly elongated and weakened. The activation barrier height h referenced to the CH4 complex is ~9 and ~19 kcal/mol for the singlet and triplet, respectively, and that referenced to the primary products Al12Ti(CH3)(H) is ~21 kcal/mol. The barrier to migration of the CH3 group around the metal cluster is estimated at ~10 kcal/mol. At the initial stage of the reaction Al12Ti + C2H6, two types of C2H6 · Al12Ti adsorption complexes are formed. In one of them, the ethane molecule is coordinated through a methyl group (as the methane molecule); and in the other type, the coordination is through the C-C bond. This reaction can proceed through two paths by means of insertion into C-H or C-C bonds to give Al12Ti(C2H5)(H) or Al12Ti(CH3)2, respectively. The second path is impeded by a high barrier (~30 kcal/mol) and is possible, if at all, only at high temperatures. Conversely, the insertion into a C-H bond in ethane is somewhat more favorable than in methane. Analogously, the PES of addition of the second methane molecule to Al12Ti(CH3)(H) was calculated. The second molecule is adsorbed and dissociates by the same mechanism as the first CH4 molecule, but with somewhat lower barriers and energy effect of formation of Al12Ti(CH3)2(H)2. The addition of propane and longer hydrocarbons is briefly considered. The results are compared with the results of previous analogous calculations of the PES of related reactions of dissociative adsorption of dihydrogen on the Al12Ti cluster, which are more exothermic, have lower barriers, and can occur under milder conditions.  相似文献   

3.
These contributions explore interaction modes between the methanethoil (CH3SH) molecule and the Fe(100) surface via implementing accurate density functional theory (DFT) calculations with the inclusion of van der Waals corrections. We consider three adsorption sites over the Fe(100) surface, namely, top(T), bridge (B), and hollow (H) sites as potential catalytic active sites for the molecular and dissociative adsorption of the CH3SH molecule. The molecular adsorption structures are found to occupy either B or T sites with former sites holding higher stability by 0.17 eV. The inclusion of van der Waals corrections refound to slightly alter adsorption energies. For instance, adsorption energies increased by ~ 0.18 and ~ 0.21 eV for B and T structure, respectively, in reference to values obtained by the plain generalized gradient approximation (GGA) functional. A stability ordering of the dissociation products was found to follow the sequence (CH4, S) > (CH3, S, H) > (─SCH3, H) > (─CH3, SH). The differential charge density distributions were examined to underpin prominent electronic contributing factors. Direct fission of C─S bond in the CH3SH molecule attains exothermic values in the range 2.0 to 2.1 eV. The most energetically favorable sites for the surface-mediated fission of the thiol's S─H bond correspond to the structure where the ─SCH3 and H are both situated on hollow sites with an adsorption energy of −2.43 eV. Overall, we found that inclusion of van der Waals functional to change the binding energies more noticeably in case of dissociative adsorption structures. The results presented herein should be instrumental in efforts that aim to design stand-alone Fe desulfurization catalysts.  相似文献   

4.
In this work, the potential energy curves of several low-lying excited states of M+(H2O)n = 1-4 (M = Li and Na) clusters with one M─O bond, related to the stretching of their M─O bond, were calculated in the gas phase. The time-dependent density functional theory and direct-symmetry-adapted cluster-configuration interaction were used in this study separately. Theoretical calculations showed that the charge transfer occurred between M+ and (H2O)n in the excited clusters so that the neutral metal atom was obtained at the dissociation limit of the potential curves. The excited potential curves of clusters were also calculated in the presence of the electrostatic field of water (EFW), and it was found that the charge transfer was blocked in the presence of EFW. The effect of the size of the (H2O)n cluster on the shape of the excited potential curves was investigated to observe how the M─O bond was affected in the excited states depending on the (H2O)n size. It was found that the increase in the size of the (H2O)n cluster increased the number of bonding excited potential curves. The difference between the electron density of the excited and ground electronic states was calculated to see how the charge transfer was affected by the size of the (H2O)n cluster.  相似文献   

5.
The thermal gas‐phase reactions of the closed‐shell metal‐oxide cluster [Al2TaO5]+ with methane have been explored by using FT‐ICR mass spectrometry complemented by high‐level quantum chemical calculations. Mechanistic aspects have been addressed to reveal the origins of the efficient addition process which results in activating the C?H bond of methane. The [Al2TaO5]+/CH4 couple has been compared with several other systems reported previously, and the electronic origins of their rather distinct performances are discussed.  相似文献   

6.
The conical intersections corresponding to the C─O and C─C ring opening were optimized and the reaction paths traversing these intersections were obtained. Investigation of the C─O ring opening revealed that when traversing the lowest energy conical intersection, the reaction path returns to the closed ring geometry. The C─O path traversing the intersection featuring torsion of terminal CH2 group however, led to a ring-opened geometry, an H-shift and the formation of acetaldehyde that can undergo further dissociation. The observation of different reaction paths was explained by the 3-D paths from quantum theory of atoms in molecules (QTAIM) that defined the most preferred direction of electronic motion that precisely tracked the mechanisms of bond breaking and formation throughout the photo-reactions. The size, orientation, and location of these most preferred 3-D paths indicated the extent and direction of motion of atoms, bonds, and the degree of torsion or planarity of a bond indicating a predictive ability.  相似文献   

7.
The electronic structures of small Al n ,n=5, 9, 13, clusters with bulk geometry are studied using the ab initio Hartree-Fock-LCAO method. The cluster ground states have always multiplicity higher than the lowest possible value. However, the energy difference between ground and lowest low spin state decreases with increasing cluster size. The energy range of the Al n cluster valence levels is comparable with the width of the occupied part of the 3sp band in bulk Al. The different binding mechanisms that arise when a CO molecule interacts with Al n clusters in different coordination sites are analyzed in detail with the constrained space orbital variation (CSOV) method. Electrostatic and polarization contributions to the interaction are found to be important. Among charge transfer (donation) contributions π electron transfer from Al n to CO corresponding to π backbonding is energetically more important than σ electron transfer from CO to Al n characterizing the σ bond.  相似文献   

8.
The structural, electronic, energy, and vibrational characteristics of the Al13X? and Al13X 2 ? clusters, with an aluminum-centered (Alc) icosahedral cage Al13 and with one or two outer-sphere ligands X=H, F, Cl, Br, OH, NH2, CH3, C6H5, have been calculated within the B3LYP approximation of the density functional theory using the 6-31G* and 6-311+G* basis sets. In all Al13X? radicals, the unpaired electron is localized at the cage atom Al* located opposite the Al-X bond. This Al* atom is the most favorable site for attaching the second X ligand of any nature (trans-addition rule). According to the previously suggested molecular model of the valence state of the [Al 13 ? ] “superatom,” the calculated energies D 1(Al 13 ? -X) of addition of the first ligand to the Al 13 ? anion are about 1 eV lower than the corresponding energies of addition of the second ligand D 2(XAl 13 ? -X). The structure of the Al13 cage depends on the nature of the nature of the substituent X and can radically change in going from anions to their neutral congeners. In the lowest-lying Al13X isomer with electronegative substituents X (Hal, OH, NH2, CH3, etc.), the aluminum cage has a marquee structure (1, symmetry C s) with a hexagonal base and a pentagonal “roof.” For Al13X analogues with electropositive ligands X (Al, Li, Na), a tridentate isomer (T, C 3v ) with the X substituent coordinated to a face of the Al13 icosahedron is preferable. In the case of moderately electronegative X ligands (of the H type), the marquee (1) and icosahedral (T) isomers are close in energy. The stretching vibration frequencies of isomers 1 and T differ significantly in magnitude and intensity so that vibrational spectroscopy methods can be especially applicable to their experimental identification.  相似文献   

9.
1,2-Eliminations are a varied and extensive set of dissociations of ions in the gas phase. To understand better such dissociations, elimination of CH2=CH2 and CH3CH3 from (CH3)2NH+CH2CH3 (1) and of CH4 from (CH3)2NH2+ are characterized by quantum chemical calculations. Stretching of the CN bond to ethyl is followed by shift of an H from methyl to the bridging position in ethyl and then to N to reach (CH3)2NH2+ + CH2=CH2 from 1. CH3CH3 elimination by H-transfer to C2H5+ to form CH3NH+=CH2 + CH3CH3 also takes place. (CH3)2NH2+ eliminates methane by CN bond extension followed by β-H-transfer to give CH2=NH+ + CH4. Low-energy reactions resembling complex-mediated 1,2-eliminations occur and constitute a hitherto largely unrecognized type of reaction. As in many complex-mediated reactions, these reactions transfer H between incipient fragments. They are distinguished from complex-mediated processes by the fragments not being able to rotate freely relative to each other near the transition state for reaction, as they do in complexes. Most 1,2-eliminations are ion-neutral complex-mediated, occur by the just described lower energy reactions, have 1,1-like transition states, or utilize highly asynchronous 1,2 transition states. All of these avoid synchronized 1,2-transition states that would violate conservation of orbital symmetry.  相似文献   

10.
《Chemical physics》1986,105(3):397-416
We report the results of a comprehensive investigation of the two-color threshold photoionization of jet-cooled indole clusters. Using two-color photoionization spectroscopy, we have probed both the neutral excited levels and the ground ionic states of indole clusters containing non-polar (Ar, CH4, CF4, C6H6) and polar (H2O, MeOH, EtOH, NH3, N(CH3)3) solvent species. These studies have allowed the determination of accurate cluster ionization energies (IEs) as well as the assignment of electronic absorption features to clusters of known composition. The determination of the cluster IE, which is typically lower than that of bare indole, has allowed us to investigate the importance of charge-induced dipole and charge-dipole attractive forces in the binding of the ion-neutral clusters. In addition, we have found that the shape of the photoionization efficiency (PIE) spectra gives valuable information regarding the relative shape and/or position of the potential energy surfaces of the neutral excited and ground ionic states of the clusters. We have also identified two distinct conformational isomers of the indole-(H2O)1 hydrogen bonded cluster using the techniques of electronic spectroscopy, two-color threshold ionization spectroscopy and mass analysis.  相似文献   

11.
Reaction energy profiles for [Me2AlNH2]3 have been computationally explored by using density functional theory. Both intra‐ and intermolecular methane elimination reactions, as well as Al?N bond‐breaking pathways, were considered. The results show that the energy required for Al?N bond breaking in cyclic [Me2AlNH2]3 is of the same order of magnitude as the activation energies for the first (limiting) step of methane elimination (for both mono‐ and bimolecular mechanisms). Thus, dissociative and associative reaction pathways are competitive. Low‐temperature/high‐pressure conditions will favor the bimolecular pathway, whereas at high temperatures, either intramolecular methane elimination or Al?N bond‐breaking dissociative pathways will be operational.  相似文献   

12.
The electronic states of small Al n (n=2–8) clusters have been calculated with a relativistic ab-initio MO-LCAO Dirac-Fock-Slater method using numerical atomic DFS wave-functions. The excitation energies were obtained from a ground state calculation of neutral clusters, and in addition from negative clusters charged by half an electron in order to account for part of the relaxation. These energies are compared with experimental photo-electron spectra.  相似文献   

13.
The structures, stabilities and the isomerization reactions of CH3SO2 isomers in a doublet electronic state have been studied at B3LYP/6‐311+ +G (d,p), MP2/6‐311++G (d,p) and CCSD(T)/6‐311++G (d,p) levels. The three different levels of calculation give the similar results: thirteen minimum isomers were located and they were connected by eleven transition states. Among the thirteen isomers, cis‐CH3OSO, trans‐CH3OSO and CH3SO2 are the most stable species, and they should be detected easily in experiment. This is well consistent with the experimental result. These isomers could isomerize to each other by chemical bond vibration, chemical bond rotation and atom migration. The non‐planar ring structure transition state (STS), which was found in this paper, extended the concept of ring STS to the non‐planar systems.  相似文献   

14.
Dissociation processes of the organoaluminum compounds Al2(CH3)6 and Al2(CH3)3Cl3 have been studied in the range of valence and Al:2p core-level ionization by means of photoelectron–photoion and photoion–photoion coincidence techniques. The double-ionization threshold and the Al:2p core-ionization threshold of Al2(CH3)6 are estimated to be about 30 and 80 eV
  • 1 1 eV = 96.4853 kJ mol?1.
  • respectively. The relative yields of the H+?Al+ and H+?CHm,+ (m′ = 0–3) ion pairs are enhanced around the Al:2p core-ionization threshold of Al2(CH3)6. The photoion–photoion coincidence intensities of Al2(CH3)3Cl3 are negligibly small throughout the energy range studied. The ratio of the relative yield of AlC2H6+ to that of Al+ increases smoothly through the Al:2p core-ionization and/or excitation region of Al2(CH3)3Cl3. The variation of the fragmentation pattern with photon energy is discussed in conjunction with the relevant electronic states.  相似文献   

    15.
    Steamed zeolites exhibit improved catalytic properties for hydrocarbon activation (alkane cracking and dehydrogenation). The nature of this practically important phenomenon has remained a mystery for the last six decades and was suggested to be related to the increased strength of zeolitic Bronsted acid sites after dealumination. We now utilize state-of-the-art infrared spectroscopy measurements and prove that during steaming, aluminum oxide clusters evolve (due to hydrolysis of Al out of framework positions with the following clustering) in the zeolitic micropores with properties very similar to (nano) facets of hydroxylated transition alumina surfaces. The Bronsted acidity of the zeolite does not increase and the total number of Bronsted acid sites decreases during steaming. O5Al(VI)-OH surface sites of alumina clusters dehydroxylate at elevated temperatures to form penta-coordinate Al1O5 sites that are capable of initiating alkane cracking by breaking the first C-H bond very effectively with much lower barriers (at lower temperatures) than for protolytic C-H bond activation, with the following reaction steps catalyzed by nearby zeolitic Bronsted acid sites. This explains the underlying mechanism behind the improved alkane cracking and alkane dehydrogenation activity of steamed zeolites: heterolytic C-H bond breaking occurs on Al-O sites of aluminum oxide clusters confined in zeolitic pores. Our findings explain the origin of enhanced activity of steamed zeolites at the molecular level and provide the missing understanding of the nature of extra-framework Al species formed in steamed/dealuminated zeolites.  相似文献   

    16.
    Ammonia is an important molecule due to its wide use in the fertiliser industry. It is also used in aminolysis reactions. Theoretical studies of the reaction mechanism predict that in reactive complexes and transition states, ammonia acts as a hydrogen bond donor forming N−H⋅⋅⋅O hydrogen bond. Experimental reports of N−H⋅⋅⋅O hydrogen bond, where ammonia acts as a hydrogen bond donor are scarce. Herein, the hydrogen bond donor ability of ammonia is investigated with three chalcogen atoms i. e. O, S, and Se using matrix isolation infrared spectroscopy and electronic structure calculations. In addition, the chalcogen bond acceptor ability of ammonia has also been investigated. The hydrogen bond acceptor molecules used here are O(CH3)2, S(CH3)2, and Se(CH3)2. The formation of the 1 : 1 complex has been monitored in the N−H symmetric and anti-symmetric stretching modes of ammonia. The nature of the complex has been delineated using Atoms in Molecules analysis, Natural Bond Orbital analysis, and Energy Decomposition Analysis. This work presents the first comparison of the hydrogen bond donor ability of ammonia with O, S, and Se.  相似文献   

    17.
    18.
    The [M21+2H]2+ cluster of the zwitterion betaine, M = (CH3)3NCH2CO2, formed via electrospray ionisation (ESI), has been allowed to interact with electrons with energies ranging from >0 to 50 eV in a Fourier transform ion cyclotron resonance (FT‐ICR) mass spectrometer. The types of gas‐phase electron‐induced dissociation (EID) reactions observed are dependent on the energy of the electrons. In the low‐energy region up to 10 eV, electrons are mainly captured, forming the charge‐reduced species, {[M21+2H]+ . }*, in an excited state, which stabilises via the ejection of an H atom and one or more neutral betaines. In the higher energy region, above 12 eV, a Coulomb explosion of the multiply charged clusters is observed in highly asymmetric fission with singly charged fragments carrying away more than 70% of the parent mass. Neutral betaine evaporation is also observed in this energy region. In addition, a series of singly charged fragments appears which arise from C? X bond cleavage reactions, including decarboxylation and CH3 group transfer. These latter reactions may arise from access of electronic excited states of the precursor ions. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

    19.
    The first σ complexes were found in the 1960s and 1970s, but they did not attract more than passing attention. Only now are we beginning to recognize their key role in the chemical reactions of σ bonds, and this has encouraged more detailed study. In contrast with the more familiar π-donor complexes such as M? (CH2?CH2) and complexes like M? NH3, in which the one pair of electrons on the N atom is bound to the metal atom, in a σ complex an X? H group binds to the transition metal atom; the X? H σ-bonding electron pair acts as a 2e donor to give an (X-H)-M type complex. Dihydrogen complexes (X = H) are one important group of σ complexes. C-H-M complexes (X = R3C) with an agostic C-H-M interaction have not only been found in the ground state but also implicated in the transition states of many important organometallic transformations such as Ziegler–Natta catalysis and sigma bond metathesis. The importance of X? H bond activation will encourage continued growth in this field.  相似文献   

    20.
    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