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.
Structural characteristics, vibrational frequencies, and energies of isomers of compounds of CO and CO2 molecules with the centered aluminum cluster Al13 and its doped analogues Al12M (M = Ti and Ni) have been calculated by the density functional theory method. For the Al12MCO compounds, the most favor-able are two “fragment” isomers in which the C and O atoms are separated and built into the cluster cage, completing it to a 14-vertex polyhedron. In one of them, the C and O atoms are in the capping positions over adjacent trigonal MAl2 faces; in the second isomer, there is the five-coordinate C* atom located in the center of a tetragonal MAl3 face and bound to the central Al atom through the long fifth bond. The “coordinated” isomers, in which the CO molecule is coordinated as a ligand to a cluster vertex, edge, or face, are unstable to removal of CO for Al13CO, close in energy to the fragment isomers for Al12NiCO, and considerably higher on the energy scale than the fragment isomers but remain stable to CO removal for Al12TiCO. For the Al12MCO2 compounds, the most favorable is the fragment isomer in which both oxygen atoms are in the capping positions over adjacent faces and the C* atom is five-coordinate. The alternative oxo carbonyl isomer Al12MO(CO) is close to the lowest-lying one in the case of M = Ni and is ~56 kcal/mol higher on the energy scale in the case of M= Ti. The less stable Al12M(CO2) isomer is the complex in which the CO2 ligand is coordinated to an M-Al edge. According to calculations, addition of CO to Al12MO and addition of CO2 to Al12M to form, respectively, Al12MO(CO) and Al12M(CO2) can occur without noticeable barrier. The Al12M(CO2) and Al12MO(CO) isomers are separated by a barrier, moderate for M = Ti (~16 kcal/mol) and small for M = Ni (~6 kcal/mol).  相似文献   

2.
The potential energy surfaces (PES) of the reactions FeAl12 + Н2 → FeН2Al12 (1) and CoAl12 + Н2 → CoН2Al12 (2) of dissociative addition of an H2 molecule to Fe- and Co-doped aluminum clusters have been calculated by the density functional theory method. Local minima on the PES in the vicinity of low-lying isomers, intermediates, and transition states have been found, and their structural and spectroscopic characteristics and energies have been calculated. The energies of the successive stages of the catalytic cycle have been evaluated, and the channels corresponding to the minimum energy path of the reactions have been studied. Differences between the structural characteristics and energies of key structures in reactions (1) and (2) have been considered. The results are compared with previous calculations of the PES of hydrogenation reactions performed for related clusters doped with nickel and titanium atoms.  相似文献   

3.
The potential energy surfaces (PES) of the elementary catalytic cycle of early stages of the H2 + MAl12 reaction of dissociative addition of an H2 molecule to aluminum clusters MAl12 doped with “light” 3d and 4d transition metal atoms (Sc, Y, Ti, Zr, V, Nb) in the states of different multiplicity have been calculated by the density functional theory method. The effect of the dopant nature and the electronic state multiplicity of the cluster on the energies and activation barriers of hydrogenation reactions of aluminum clusters is considered. The calculated PES corresponding to the early stages of the H2 + TiAl12 reaction does not reveal any specific features that could be evidence of the significant preference of the titanium dopant as compared with other transitions metals like Zr or W.  相似文献   

4.
Inspired by the concept of superatom via substitutionally doping an Al13 magic cluster, we investigated the H2 molecule dissociation on the doped icosahedral Al12X (X = B, Al, C, Si, P, Mg, and Ca) clusters by means of density functional theory. The computed reaction energies and activation barriers show that the concept of superatom is still valid for the catalysis behavior of doped metal clusters. The hydrogen dissociation behavior on metal clusters characterized by the activation barrier and reaction energy can be tuned by controllable doping. Thus, doped Al12X clusters might serve as highly efficient and low‐cost catalysts for hydrogen dissociation. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

5.
Structures, energies, and spectroscopic characteristics of the isomers of the family of doubly doped М2Al42 clusters with dopants M from the first two periods and H, Cu, and Zn atoms located in different positions at the surface and in the inner cavity of the aluminum cage have been calculated by the density functional theory method. The effect of the dopant nature on the relative energies of isomers and on the energies of their dissociation along the channels М2Al42 → 2М + Al42 and М2Al42 + 2Al → 2М+ Al44. The results are compared with the results of previous DFT calculations of endohedral (M@Al12) and exohedral (Al@)MAl11) isomers of the simpler doped clusters MAl12 with the same dopants M. The influence of the aluminum cage size on the relative energy stability of the surface and and internal sopant positions is considered.  相似文献   

6.
Structural characteristics, vibrational frequencies, and energies of ten isomers of acetylene compounds with the centered aluminum cluster Al13 and its analogues Al12M doped with 3d transition metal atoms (M = Ti-Ni) in the states of different multiplicity have been calculated by the density functional theory method. In addition to “coordinated” intermediates in which the C2H2 molecule is coordinated through its C-C bond to the M vertex, an M-Al edge, or a trigonal face of the [MAl2] cluster, “fragment” isomers have been considered in which the acetylene molecule is broken into fragments (C2H + H, CH + CH, H + CH + C, and 2P + 2H) differently inserted into the aluminum cage and enlarging it to Al12MC2. For most compounds, low-lying isomers have structures 1–4 (the C2H2 molecule is coordinated to an Al2M face), 1–5 (two CH fragments are added to adjacent Al2M faces), and 1–8 (with a five-coordinate C* atom). Structure 1–1, in which the C2H2 molecule is coordinated through the C-C bond to the M dopant is unstable against transformation into 1–4 with a low barrier. An isomer with unusual structure 1–9 has been localized in which two five-coordinate C* atoms built into the aluminum cage are located in adjacent quasi-planar tetragonal [MAl3] faces and are bonded to the central aluminum atom (Alc) through the fifth bonds. The substitution of electronegative substituents X= F and Cl for H atoms in isomers 1–8 and 1–9 makes the latter more basic and clearly more favorable. The five-coordinate C* atoms in them are able to add acceptor ligands of the BH3 and AlH3 type and to increase the coordination number of the carbon atom to six with a considerable decrease in energy. The trends in the change in structural characteristics and relative energies of isomers with a change in M dopants along the 3d series, electronegativity of X substituents, and electronic state multiplicity have been analyzed.  相似文献   

7.
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   

8.
This study provides details of the electronic and optical structures and binding energies of sarin (SF) and chlorosarin (SC) with Al–N and Al–P surfaces of Al12N12 and Al12P12 nanoclusters in the gas phase. The adsorption mechanism of SF and SC on these nanoclusters containing the Al3+ central cation was studied. Optimized geometries and thermodynamic parameters of SF and SC adsorption complexes were calculated. SF and SC are chemisorbed on these nanoclusters because of the formation of PO···Al bonds. The chemical bond is formed between an oxygen atom of SF and SC and an aluminum atom of fullerene-likes (chemisorption). However, the binding energies of the complexes with the Al12N12 nanocluster are larger than these values for the Al12P12 nanocluster. The interaction enthalpy and Gibbs free energy of all studied systems were found to be negative. We can conclude that SF and SC will be adsorbed preferably on Al12N12 nanocluster.  相似文献   

9.
The Al12Ti(π-L) complexes with ligands L = C2H2, C2H4, HCN, N2H2, C6H6, and N2 in the singlet and triplet states have been calculated within the B3LYP approximation of the density functional theory using the 6–31G* basis set. Their calculated structures and properties have been compared with the results of analogous calculations of the titanium porphyrin complexes with the same ligands L. It has been demonstrated that, in both series of compounds, the side-on coordination of the ligands (through the multiple bond, π type) to the titanium atom is accompanied by the weakening and elongation of the C-C, C-N, and N-N bonds by 0.05–0.20Å and the long-wavelength shift of the stretching vibration modes νstr(CC), νstr(CN), and νstr(NN) by a few hundreds of cm?1. For the Ti aluminide complexes, these activation effects are much more clearly pronounced than for their Ti porphyrin analogues. The aluminide complexes (except the nitrogenyl one) have the singlet ground state; however, the nearest triplet is close lying to the singlet (within 1–14 kcal/mol). The singlet is characterized by the considerable electron density transfer from the Al12Ti cluster to the ligand L, the displacement of the Ti atom from the aluminum cage to the ligand, and the distortion of the Al12 cage. In the triplet states, the ligand activation depends on the character of spin density delocalization between the ligand, the Ti atom, and the Al12 cage. If the spin density is distributed between the Ti atom and the Al12 cage or if both unpaired electrons are localized on the Ti atom, then the structure, stability, and spectroscopic properties of the “active” Ti-L moiety in these triplets differ only slightly from those in the singlet state. If the spin density is distributed between the Ti atom and the ligand, the singlet-triplet excitation is accompanied by the elongation and weakening of the Ti-ligand bond and the decrease in the ligand activation effect. Complexes with several ligands L coordinated to the Ti atom in the Al12Ti cluster have been calculated. There are some trends in the change in the molecular characteristics of the Ti-ligand bonds in different series of the complexes.  相似文献   

10.
The energies and structural and spectroscopic characteristics of a series of model stepwise hydrogenated aluminum clusters Al44H n (n = 1?24) obtained by successive introduction of hydrogen atoms into various surface positions of the Al44 cluster have been calculated by the density functional theory method (B3LYP). According to these calculations, the [Al39] surface layer of the cage retains a closed “nested doll” shape with a pentaatomic inner core [Al5]. With increasing n, both the surface layer and the core tend to experience increasing asymmetric distortions. The surface is corrugated and undergoes significant axial and equatorial extensions and contractions, some of the Al?H two-center terminal bonds are transformed into threecenter hydrogen bridges, and some Al atoms are displaced from the surface layer to the outer sphere and are bound to the surface through hydrogen bridges. The inner core [Al5] at n = 24 loses its bipyramidal shape and shifts to the surface layer so that one or two of its atoms are “built-in” into the concave regions of the surface layer. The calculated average energies of Al?H bonds are within the range ~55.5 ± 2.5 kcal/mol. The averaged energies of the Al44H n → Al44Hn–2 + Н2 dissociation reactions with elimination of a hydrogen molecule are on the order of a few kilocalories per mole and are evidence of small exothermicity (or isothermicity) of these reactions. For the Al44H, Al44H2, and Al44H6 clusters as an example, the relative stabilities of isomers with terminal Al?H bonds in various nonequivalent positions of the [Al39] surface layer are compared.  相似文献   

11.
The interaction of hydrazine (N2H4) molecule with pristine and Si-doped aluminum nitride (Al12N12) nano-cage was investigated using the density functional theory calculations. The adsorption energy of N2H4 on pristine Al12N12 in different configurations was about –1.67 and –1.64 eV with slight changes in its electronic structure. The results showed that the pristine nano-cage can be used as a chemical adsorbent for toxic hydrazine in nature. Compared with very low sensitivity between N2H4 and Al12N12 nano-cage, N2H4 molecule exhibits high sensitivity toward Si-doped Al12N12 nano-cage so that the energy gap of the Si-doped Al12N12 nano-cage is changed by about 31.86% and 37.61% for different configurations in the SiAl model and by about 26.10% in the SiN model after the adsorption process. On the other hand, in comparison with the SiAl model, the adsorption energy of N2H4 on the SiN model is less than that on the SiAl model to hinder the recovery of the nano-cage. As a result, the SiN Al12N11 is anticipated to be a potential novel sensor for detecting the presence of N2H4 molecule.  相似文献   

12.
The solution chemistry of aluminum has long interested scientists due to its relevance to materials chemistry and geochemistry. The dynamic behavior of large aluminum–oxo‐hydroxo clusters, specifically [Al13O4(OH)24(H2O)12]7+ ( Al13 ), is the focus of this paper. 27Al NMR, 1H NMR, and 1H DOSY techniques were used to follow the isomerization of the ?‐Al13 in the presence of glycine and Ca2+ at 90 °C. Although the conversion of ?‐Al13 to new clusters and/or Baker–Figgis–Keggin isomers has been studied previously, new 1H NMR and 1H DOSY analyses provided information about the role of glycine, the ligated intermediates, and the mechanism of isomerization. New 1H NMR data suggest that glycine plays a critical role in the isomerization. Surprisingly, glycine does not bind to Al30 clusters, which were previously proposed as an intermediate in the isomerization. Additionally, a highly symmetric tetrahedral signal (δ=72 ppm) appeared during the isomerization process, which evidence suggests corresponds to the long‐sought α‐Al13 isomer in solution.  相似文献   

13.
Reactions of cooled, size-selected aluminum cluster ions (Aln+, n = 1–8) with oxygen have been studied at collision energies from 0.15 to 10.0 eV (center-of-mass) under single-collision conditions. With the exception of the atomic ion, all size clusters undergo exoergic reactions which result in extensive fragmentation of the metal cluster framework. Significant energy barriers are found for reaction of all clusters except the dimer. The barrier height increases with cluster size from Al3+ to Al7+, then drops for Al8+.  相似文献   

14.
The potential energy surfaces of elementary reactions of dissociative addition of one and two H2 molecules to Cr-, Mo-, or W-doped aluminide clusters MAl12 in the states of different multiplicity have been calculated by the density functional theory method. The results are compared with the previous calculations of analogous reactions involving the singlet and triplet TiAl12 cluster. The effect of the dopant nature and electronic state multiplicity on the energies and activation barriers of hydrogenation reactions is considered.  相似文献   

15.
The transition states and activation barriers h of elementary reactions of addition of the H2 molecule to aluminide clusters Al13, Al 13 ? , Al13H 2 ? , Al13H 4 ? , Si@Al12, Ge@Al12, and LiAl13 were calculated within the B3LYP approximation of the density functional theory using 6–31G* and 6–311+G* basis sets. The barriers h for all diamagnetic clusters were found to be high (~30–40 kcal/mol). The outer-sphere cation Li+ decreases while the endohedral electronegative dopants Si and Ge increase the barrier by a few kcal/mol. The hydrogenation barrier of the neural paramagnetic cluster Al13, which has free valence, decreases to ~20 kcal/mol. The addition of a hydrogen atom or a Cl2 molecule to both paramagnetic and diamagnetic aluminum clusters occurs without a barrier. The first stage of the reaction (addition of H2 to an Al-Al edge) is in all cases the critical stage of aluminide hydrogenation. The barrier h of this reaction is several times higher than the barriers to migration of hydrogen atoms over the metal cage. The migration of H atoms occurs simultaneously with considerable distortions of the Al13 cage even to the extent that it changes its structural motif. The addition of the H2 molecule to the Al@TiAl11 cluster containing the peripheral titanium atom occurs with a small barrier, whereas the barrier to elimination of H2 from the dihydride Al@TiAl11H2 is reduced to ~15 kcal/mol. Based on the calculations, the conclusion was drawn that the elementary reactions of hydrogenation and dehydrogenation for Ti-doped aluminide clusters should occur considerably faster and under milder conditions than for homonuclear aluminides.  相似文献   

16.
Density functional method has been employed to compare the interaction and adsorption abilities of simple and doped fullerenes with various heteroatoms (Al, B, Si, N, P, and S). Three sulfur‐containing molecules (H2S, SO2, and thiophene) were selected to study of their interactions with fullerenes. These interactions will be important in the design of new sensor, adsorption, and elimination of pollutants and chemical reactions. The calculated adsorption energies (Ead) in the gas phase and solvents (water, using the polarized continuum model) showed that all adsorbates have exothermic interaction with all fullerenes. The maximum Ead values were calculated for aluminum‐doped fullerene (AF) and nitrogen‐doped fullerene (NF), and the adsorption energies in solvent are not so different with those in the gas phase. Natural Bond Orbitals (NBO) calculations showed the complexes of AF and then boron‐doped fullerene (BF) have the highest E2 interaction energies, whereas simple fullerene (F) and phosphorus‐doped fullerene (PF) have the least E2 energies. Population analyses showed that doping by heteroatoms bearing extra electrons reduces the energy gap and this decrease is more than the decrease observed from doping by heteroatoms with electron defect. Moreover, the change in the energy gaps of the complexes, obtained from the density of states (DOSs) plots, showed that these structures could be used in sensor devices. All calculated data confirmed the better adsorption of SO2 by fullerenes versus H2S and thiophene and among all fullerenes, AF and then BF and NF are the best adsorbent for these structures.  相似文献   

17.
More than 20 М6Al38 isomers and several М12Al32 isomers for nitrogen- and phosphorus-substituted clusters with six and twelve dopant atoms M = N and P substituted for Al atoms in different positions at the surface of the aluminum cage and inside it have been studied by the density functional theory method. In the preferred N6Al38 isomer, all N atoms are substituted for Al atoms initially located in one outer layer of the cluster. In the course of geometry optimization, the nitrogen atoms are incorporated into positions in the neighboring intermediate layer, thus converting it into a 12-atom face consisting of three vertex-sharing adjacent six-membered rings with short N–Al bonds. For Р6Al38, a distribution of the dopant either in both surface layers or in the intermediate space between the surface layers and the inner core of the cluster is preferred. Optimization of alternative structures of the N12Al32 cluster with N atoms substituted for Al atoms in both outer layers is evidence in favor of the isomer in which the dopants are dispersed as separated monatomic anions N–. Together with their bridging Al atoms, these anions form the inner [N12Al14] cage with an unusual dumbbell-like structure in which the upper and lower halves are linked through N–Al bonds with the equatorial aluminum atoms. In the next low-lying isomer being ~23 kcal/mol higher on the energy scale, there is observed the “microclustering” of the dopant to form three covalently bonded diatomic dianions N22-; the latter, together with the bridging Al atoms are combined into a [N6Al6] “subcluster” inside the severely distorted outer cage. In P12Al32, the aluminum cage is subjected only to moderate distortions: the phosphorus atoms remain in the outer layers and form two three-membered rings [Р3]. The estimated energies of the model substitution reactions Al44 + M6 → M6Al38 + Al6 (1) and Al44 + 2M6 → M12Al38 + 2Al6 (2) demonstrate that all these reactions are exothermic; however, for the nitrogen-containing clusters, the decrease in energy with increasing number of substitutions increases from 66 (1) to 113 (2) kcal/mol, while in the case of phosphorus, it decreases from 45 (1) to 4 (2) kcal/mol. The results obtained for N6Al38, N12Al32, Р6Al38, and Р12Al32 are compared with the previous calculations for the C6Al38, C12Al32, Si6Al38, and Si12Al32 clusters.  相似文献   

18.
More than twenty M6Al38 isomers and several M12Al32 isomers of carbon- and silicon-substituted aluminum clusters with six and twelve dopant atoms of general formula MnAl44–n(M = C and Si, n = 6 and 12) have been studied by the density functional theory method. Calculations predict that, in the lowest-lying M6Al38, isomer, all substitutions of C atoms for Al are localized in one outer surface layer of the aluminum cage. In the course of optimization, the C atoms with a negative charge of about 1e are incorporated into positions of the intermediate layer to transform it into a 12-atom face composed of three adjacent vertex-sharing six-membered rings with short C–Al bonds. In the favorable isomer of M6Al38, the dopants are scattered as individual Si atoms located in both outer layers or in the subsurface space between the outer layers and the inner core of the cluster. Optimization of low-lying isomers with twelve starting substitutions of C and Si for Al in both outer layers has localized two preferable C12Al32 isomers. One of them contains three covalently bonded diatomic C2 anions, which are combined through bridging aluminum atoms in the three-dimensional [C6Al7] cluster inside the severely distorted outer cage. In the second, most favorable, isomer, the dopants are distributed as isolated C anions; together with the bridging Al atoms, they form the [M12Al32] inner cage with an unusual dumbbell-like structure. For M12Al32, the aluminum cage undergoes moderate distortions. The silicon atoms remain in the outer layers and form five-membered ring subclusters [Si5] and [Si2Al3] bound to the neighboring intermediate layers through elongated and weakened Si–Al bonds. Evaluation of the energies of the model exchange reactions Al44 + M6 → M6Al38 + Al6 and Al44 + 2M6 → M12Al32 + 2Al6 shows that for M= C both reaction are exothermic, whereas for M = Si the former reaction is nearly isothermal and the second reaction is endothermic and requires significant energy inputs. The differences between the equilibrium structures and the relative positions on the energy scale of the isomers of the C6Al38–Si6Al38 and C12Al38–Si12Al38 clusters are examined.  相似文献   

19.
Heptene-2 aromatization on Pt/Al2O3 in a pulse microcatalytic reactor has been studied under H2 and N2 atmosphere at temperatures between 330 to 500°C and at a total pressure of 4.0 kg/cm2. Results showed that the production of only cracked products (mainly methane) from deep fragmentation of heptene-2 in H2 sharply contrasts with the reaction in N2 in which the catalyst showed aromatic selectivity with the production of methane, benzene and toluene. In H2-N2 mixtures, 75% H2 was required to reduce the aromatization activity of the catalyst to zero. A test of the kinetic data using Sica's method [15] of pulse kinetic analysis suggests a first order in heptene-2 with an activation energy of 102.61 kJ/mol in N2 and 124.71 kJ/mol in H2. The difference in activation energies has been attributed to a difference in reaction mechanisms in both gases.  相似文献   

20.
The electronic and geometric structures of different isomers of the closo-B10H12 boron cluster have been calculated by the density functional theory method (in the B3LYP/6-311++G**//B3LYP/6-31G* approximation). The compound is considered to be the diprotonated (H*) analogue of the well-studied B10H 10 2? anion and serves as a model system. The increase in the relative energies of isomers and the preferred location of the extra H* protons near the opposite B(1) and B(10) “poles” are consistent with the charge separation (in the framework of the Mulliken population analysis) between B(1) and B(10). The reactions of migration of one or simultaneously two H* protons in B10H12 over the boron polyhedron have been considered, and the corresponding energies of elementary events E and activation barriers h have been estimated. The elementary events have been predicted in which both H* protons simultaneously move along the trajectories near the opposite B(1) and B(10) poles of the B10H 10 2? polyhedron with the same or opposite changes in the angles determining the H* position with respect to the B(1)–B(10) axis. The activation barrier to the “opposite” migration of the H* protons has been assessed to be h ~ 1.2 kcal/mol, whereas for the migration of the H* protons in the same direction, h ~ 1.4 kcal/mol. The H* proton transfer from the position near the B(1) pole to the position near the opposite B(10) pole is hindered, and higher activation barriers on the order of h ~ 13–15 kcal/mol should be overcome for this transfer to occur.  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