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 15 毫秒
1.
The energetics of the Menshutkin-like reaction between four mesylate derivatives and ammonia have been computed using B3LYP functional with the 6-31+G** basis set. Additionally, MPW1K/6-31+G** level calculations were carried out to estimate activation barrier heights in the gas phase. Solvent effect corrections were computed using PCM/B3LYP/6-31+G** level. The conversion of the reactant complexes into ion pairs is accompanied by a strong energy decrease in the gas phase and in all solvents. The ion pairs are stabilized with two strong hydrogen bonds in the gas phase. The bifurcation at C2 causes a significant activation barrier increase. Also, bifurcation at C5 leads to noticeable barrier height differentiation. Both B3LYP/6-31+G** and MPW1K/6-31+G** activation barriers suggest the reaction 2 (2a?+?NH3) to be the fastest in the gas phase. The reaction 4 is the slowest one in all environments.
Figure
Ammonium salt formation in a Menshutkin-like reaction between ammonia and (S)-1,4-andydro-2,3-dideoxy-5-O-mesylpentitol (2a)  相似文献   

2.
Density functional theory (DFT) calculations at B3LYP/6-31 G (d,p) and B3LYP/6-311?+?G(d,p) levels for the substituted pyridine-catalyzed isomerization of monomethyl maleate revealed that isomerization proceeds via four steps, with the rate-limiting step being proton transfer from the substituted pyridinium ion to the C=C double bond in INT1. In addition, it was found that the isomerization rate (maleate to fumarate) is solvent dependent. Polar solvents, such as water, tend to accelerate the isomerization rate, whereas apolar solvents, such as chloroform, act to slow down the reaction. A linear correlation was obtained between the isomerization activation energy and the dielectric constant of the solvent. Furthermore, linearity was achieved when the activation energy was plotted against the pK a value of the catalyst. Substituted-pyridine derivatives with high pK a values were able to catalyze isomerization more efficiently than those with low pK a values. The calculated relative rates for prodrugs 16 were: 1 (406.7), 2 (7.6?×?106), 3 (1.0), 4 (20.7), 5 (13.5) and 6 (2.2?×?103). This result indicates that isomerizations of prodrugs 1 and 35 are expected to be slow and that of prodrugs 2 and 6 are expected to be relatively fast. Hence, prodrugs 2 and 35 have the potential to be utilized as prodrugs for the slow release of monomethylfumarate in the treatment of psoriasis and multiple sclerosis.
Figure
Substituted pyridine-catalyzed isomerization of monomethylmaleate (prodrug, cis-isomer) to monomethylfumerate (parental drug, trans-isomer)  相似文献   

3.
In this work we have performed a systematic study of new organometallic complexes containing penta- and heteropentadienyl (CH2CHCHCHX, X?=?CH2, O, NH, S) ligands coordinated to beryllium. Calculated complexes were studied using the density functional theory (PBE) in combination with the 6-311++G(3d,2p) basis set. The coordination number on the beryllium atom varies according to the type of ligand. Pentadienyl ligand shows hapticities η1 and η5, while heteropentadienyl ligands display η1 and η2 hapticities. A Wiberg bond indices study was performed in order to get information about their bond orders.
Figure
Organometallic structures with beryllium and heteropentadienyl ligands  相似文献   

4.
Density functional theory (DFT) was used to investigate the nickel- or nickel(0)/zinc- catalyzed decarbonylative addition of phthalic anhydrides to alkynes. All intermediates and transition states were optimized completely at the B3LYP/6-31+G(d,p) level. Calculated results indicated that the decarbonylative addition of phthalic anhydrides to alkynes was exergonic, and the total free energy released was ?87.6 kJ mol?1. In the five-coordinated complexes M4a and M4b, the insertion reaction of alkynes into the Ni-C bond occurred prior to that into the Ni-O bond. The nickel(0)/zinc-catalyzed decarbonylative addition was much more dominant than the nickel-catalyzed one in whole catalytic decarbonylative addition. The reaction channel CAM1'T1'M2'T2'M3a'M4a'T3a1'M5a1'T4a1'M6a'P was the most favorable among all reaction pathways of the nickel- or nickel(0)/zinc- catalyzed decarbonylative addition of phthalic anhydrides to alkynes. And the alkyne insertion reaction was the rate-determining step for this channel. The additive ZnCl2 had a significant effect, and it might change greatly the electron and geometry structures of those intermediates and transition states. On the whole, the solvent effect decreased the free energy barriers.
Figure
DFT study suggests that NiL4/ZnCl2 (L=PMe3) has higher catalysis than NiL4 in the synthesis of isocoumarin from phthalic anhydrides and alkynes.  相似文献   

5.
The geometries, energies, and electronic properties of the two possible configurations of bis-[dibenzo[a.i]fluorenylidene] were investigated theoretically by density functional theory DFT B3LYP at the UB3LYP/6-311?+?G(2d,p) // UB3LYP/6-31?+?G(d,p) level of theory. According to the performed calculations, it was found that the singlet is 3.4?kcal?mol-1 lower in energy compared to triplet state at room temperature. This gap is compared with those of other alkenes like ethylene, (61.9?kcal?mol-1) tetra-tert-butyethylene, (6.4?kcal?mol-1) and bis-fluorenylidene (19.5?kcal?mol-1). These results confirm the experimental findings of the paramagnetic properties determined by Franzen and Joschek. The low singlet-triplet gap in the case of bis-[dibenzo[a.i]fluorenylidene] is the result of a steric destabilization of the singlet due to strain and stabilization of the triplet electronic state by delocalization of each free electron within each aromatic moiety. This correlates with the special electronic structure of the triplet state of this compound, where facial interaction of two hydrogen atoms lying close to the lobes of each p-orbital occupied with a single electron at the distorted double bond in the triplet electronic state.
Figure
a) The singlet form of bis-dibenzo[a.i]fluorenylidene. b) The triplet form of bis-dibenzo[a.i]fluorenylidene. The central dihedral angle around the C=C double bond changes from 53.2° in the singlet electronic structure to 90.0° in the triplet electronic structure. Of great interest is the very low singlet-triplet gap of this electronic system which equals to 3.4 kcal/mol according to calculation by DFT UB3LYP/6-311+G(2d,p) // UB3LYP/6-31+G(d,p) level of theory.  相似文献   

6.
The changes of bond dissociation energy (BDE) in the C–NO2 bond and nitro group charge upon the formation of the molecule-cation interaction between Na+ and the nitro group of 14 kinds of nitrotriazoles or methyl derivatives were investigated using the B3LYP and MP2(full) methods with the 6-311++G**, 6-311++G(2df,2p) and aug-cc-pVTZ basis sets. The strength of the C–NO2 bond was enhanced in comparison with that in the isolated nitrotriazole molecule upon the formation of molecule-cation interaction. The increment of the C–NO2 bond dissociation energy (ΔBDE) correlated well with the molecule-cation interaction energy. Electron density shifts analysis showed that the electron density shifted toward the C-NO2 bond upon complex formation, leading to the strengthened C-NO2 bond and the possibly reduced explosive sensitivity.
Figure
C1-N2 bond turns strong upon molecule-cation interaction formation, leading to a possibly reduced explosive sensitivity.  相似文献   

7.
Quantum chemical calculations were performed to investigate the stability of the ternary complexes BeH2···XMH3···NH3 (X?=?F, Cl, and Br; M?=?C, Si, and Ge) and the corresponding binary complexes at the atomic level. Our results reveal that the stability of the XMH3···BeH2 complexes is mainly due to both a strong beryllium bond and a weak tetrel–hydride interaction, while the XMH3···NH3 complexes are stabilized by a tetrel bond. The beryllium bond with a halogen atom as the electron donor has many features in common with a beryllium bond with an O or N atom as the electron donor, although they do exhibit some different characteristics. The stability of the XMH3···NH3 complex is dominated by the electrostatic interaction, while the orbital interaction also makes an important contribution. Interestingly, as the identities of the X and M atoms are varied, the strength of the tetrel bond fluctuates in an irregular manner, which can explained by changes in electrostatic potentials and orbital interactions. In the ternary systems, both the beryllium bond and the tetrel bond are enhanced, which is mainly ascribed to increased electrostatic potentials on the corresponding atoms and charge transfer. In particular, when compared to the strengths of the tetrel and beryllium bonds in the binary systems, in the ternary systems the tetrel bond is enhanced to a greater degree than the beryllium bond.
Graphical Abstract A tetrel bond can be strengthened greatly by a beryllium bond
  相似文献   

8.
The changes of bond dissociation energy (BDE) in the C–NO2 bond and nitro group charge upon the formation of the intermolecular hydrogen-bonding interaction between HF and the nitro group of 14 kinds of nitrotriazoles or methyl derivatives were investigated using the B3LYP and MP2(full) methods with the 6-311++G**, 6-311++G(2df,2p) and aug-cc-pVTZ basis sets. The strength of the C–NO2 bond was enhanced and the charge of nitro group turned more negative in complex in comparison with those in isolated nitrotriazole molecule. The increment of the C–NO2 bond dissociation energies correlated well with the intermolecular H-bonding interaction energies. Electron density shifts analyses showed that the electron density shifted toward the C–NO2 bond upon complex formation, leading to the strengthened C–NO2 bond and the possibly reduced explosive sensitivity.
Figure
C1-N2 bond turns strong upon H-bond formation, leading to a possibly reduced explosive sensitivity  相似文献   

9.
Dipole moments (μ), charge distributions, and static electronic first-order hyperpolarizabilities (β μ ) of the two lowest-energy keto tautomers of guanine (7H and 9H) were determined in the gas phase using Hartree–Fock, Møller–Plesset perturbation theory (MP2 and MP4), and DFT (PBE1PBE, B97-1, B3LYP, CAM-B3LYP) methods with Dunning’s correlation-consistent aug-cc-pVDZ and d-aug-cc-pVDZ basis sets. The most stable isomer 7H exhibits a μ value smaller than that of the 9H form by a factor of ca. 3.5. The β μ value of the 9H tautomer is strongly dependent on the computational method employed, as it dramatically influences the β μ (9H)/β μ (7H) ratio, which at the highest correlated MP4/aug-cc-pVDZ level is predicted to be ca. 5. The Coulomb-attenuating hybrid exchange-correlation CAM-B3LYP method is superior to the conventional PBE1PBE, B3LYP, and B97-1 functionals in predicting the β μ values. Differences between the largest diagonal hyperpolarizability components were clarified through hyperpolarizability density analyses. Dipole moment and first-order hyperpolarizability are molecular properties that are potentially useful for distinguishing the 7H from the 9H tautomer.
Figure
Hyperpolarizability density analysis of the most stable guanine tautomer  相似文献   

10.
11.
MP2(full)/6-311++G(3df,3pd) calculations were carried out on complexes linked through various non-covalent Lewis acid – Lewis base interactions. These are: hydrogen bond, dihydrogen bond, hydride bond and halogen bond. The quantum theory of ´atoms in molecules´ (QTAIM) as well as the natural bond orbitals (NBO) method were applied to analyze properties of these interactions. It was found that for the A-H…B hydrogen bond as well as for the A-X…B halogen bond (X designates halogen) the complex formation leads to the increase of s-character in the A-atom hybrid orbital aimed toward the H or X atom. In opposite, for the A…H-B hydride bond, where the H-atom possesses negative charge, the decrease of s-character in the B-atom orbital is observed. All these changes connected with the redistribution of the electron charge being the effect of the complex formation are in line with Bent´s rule. The numerous correlations between energetic, geometrical, NBO and QTAIM parameters were also found.
Figure
QTAIM atomic radii for NH4 +…HMgH and Na+…HBeH  相似文献   

12.
The gas phase molecular structure of a single isolated molecule of [Ag(Etnic)2NO3];1 where Etnic = Ethylnicotinate was calculated using B3LYP method. The H-bonding interaction between 1 with one (complex 2) and two (complex 3) water molecules together with the dimeric formula [Ag(Etnic)2NO3]2;4 and the tetrameric formula [Ag(Etnic)2NO3]4;5 were calculated using the same level of theory to model the effect of intermolecular interactions and molecular packing on the molecular structure of the titled complex. The H-bond dissociation energies of complexes 2 and 3 were calculated to be in the range of 12.220–14.253 and 30.106–31.055 kcal?mol?1, respectively, indicating the formation of relatively strong H-bonds between 1 and water molecules. The calculations predict bidentate nitrate ligand in the case of 1 and 2, leading to distorted tetrahedral geometry around the silver ion with longer Ag–O distances in case of 2 compared to 1, while 3 has a unidentate nitrate ligand leading to a distorted trigonal planar geometry. The packing of two [Ag(Etnic)2NO3] complex units; 4 does not affect the molecular geometry around Ag(I) ion compared to 1. In the case of 5, the two asymmetric units of the formula [Ag(Etnic)2NO3] differ in the bonding mode of the nitrate group, where the geometry around the silver ion is distorted tetrahedral in one unit and trigonal planar in the other. The calculations predicted almost no change in the charge densities at the different atomic sites except at the sites involved in the C–H?O interactions as well as at the coordinated nitrogen of the pyridine ring.
Figure
Molecular structure (left) and electrostatic potentials mapped on the electron density surface (right) calculated by DFT/B3LYP method for Etnic, and complexes 1 and 2  相似文献   

13.
DFT calculations at B3LYP/6-31G(d,p) for intramolecular proton transfer in Kirby’s enzyme models 17 demonstrated that the reaction rate is dependent on the distance between the two reacting centers, rGM, and the hydrogen bonding angle, α, and the rate of the reaction is linearly correlated with rGM and α. Based on these calculation results three simvastatin prodrugs were designed with the potential to provide simvastatin with higher bioavailability. For example, based on the calculated log EM for the three proposed prodrugs, the interconversion of simvastatin prodrug ProD 3 to simvastatin is predicted to be about 10 times faster than that of either simvastatin prodrug ProD 1 or simvastatin ProD 2. Hence, the rate by which the prodrug releases the statin drug can be determined according to the structural features of the promoiety (Kirby’s enzyme model).
Figure
A representation Scheme showing the interconversion of simvastatin prodrug to simvastatin by a prodrug chemical approach.  相似文献   

14.
The geometric and electronic structures, absorption spectra, transporting properties, chemical reactivity indices and electrostatic potentials of the planar three-coordinate organoboron compounds 1-2 and twisted reference compound Mes 3 B, have been investigated by employing density functional theory (DFT) and conceptual DFT methods to shed light on the planarity effects on the photophysical properties and the chemical reactivity. The results show that the planar compounds 1-2 exhibit significantly lower HOMO level than Mes 3 B, owing to the stronger electronic induction effect of boron centers. This feature conspicuously induces a blue shifted absorption for 1, although 1 seemingly possesses more extended conjugation framework than Mes 3 B. Importantly, the reactivity strength of the boron atoms in 1-2 is much lower than that in Mes 3 B, despite the fact that the tri-coordinate boron centers of 1-2 are completely naked. The interesting and abnormal phenomenon is caused by the strong p-π electronic interactions, that is, the empty p-orbital of boron center is partly filled by π-electron of the neighbor carbon atoms in 1-2, which are confirmed by the analysis of Laplacian of the electron density and natural bond orbitals. Furthermore, the negative electrostatic potentials of the boron centers in 1-2 also interpret that they are not the most preferred sites for incoming nucleophiles. Moreover, it is also found that the planar compounds 1-2 can act as promising electron transporting materials since the internal reorganization energies for electron are really small.
Figure
The planar effects significantly affect the frontier molecular orbital levels, absorption wavelengths, transporting properties, and chemical reactivities of compounds 1-2. The underlying origin has been revealed by density functional theory and conceptual density functional theory calculations  相似文献   

15.
16.
In the present study we have characterized the halogen bonding in selected molecules H3N–ICF3 (1-NH 3 ), (PH3)2C–ICF3 (1-CPH 3 ), C3H7Br–(IN2H2C3)2C6H4 (2-Br), H2–(IN2H2C3)2C6H4 (2-H 2 ) and Cl–(IC6F5)2C7H10N2O5 (3-Cl), containing from one halogen bond (1-NH 3 , 1-CPH 3 ) up to four connections in 3-Cl (the two Cl–HN and two Cl–I), based on recently proposed ETS-NOCV analysis. It was found based on the NOCV-deformation density components that the halogen bonding C–X B (X-halogen atom, B-Lewis base), contains a large degree of covalent contribution (the charge transfer to X B inter-atomic region) supported further by the electron donation from base atom B to the empty σ*(C–X) orbital. Such charge transfers can be of similar importance compared to the electrostatic stabilization. Further, the covalent part of halogen bonding is due to the presence of σ-hole at outer part of halogen atom (X). ETS-NOCV approach allowed to visualize formation of the σ-hole at iodine atom of CF3I molecule. It has also been demonstrated that strongly electrophilic halogen bond donor, [C6H4(C3H2N2I)2][OTf]2, can activate chemically inert isopropyl bromide (2-Br) moiety via formation of Br–I bonding and bind the hydrogen molecule (2-H 2 ). Finally, ETS-NOCV analysis performed for 3-Cl leads to the conclusion that, in terms of the orbital-interaction component, the strength of halogen (Cl–I) bond is roughly three times more important than the hydrogen bonding (Cl–HN).
Figure
ETS-NOCV reprezentation of σ-hole at iodine together with the molecular electrostatic potential picture  相似文献   

17.
The structure and electronic properties of the complexes formed by the interaction of imidazole and pyrazole with different BeXH(BeX2) (X = H, Me, F, Cl) derivatives have been investigated via B3LYP/6?311+G(3df,2p)//B3LYP/6?31+G(d,p) calculations. The formation of these azole:BeXH(BeX2) complexes is accompanied by a dramatic enhancement of the intrinsic acidity of the azole, as the deprotonated azole is much more stable after the aforementioned interaction. Most importantly, the increase in acidity is so large that the azole:BeXH or azole:BeX2 complexes behave as NH acids, which are stronger than typical oxyacids such as phosphoric acid and oxalic acid. Interestingly, the increase in acidity can be tuned through appropriate selection of the substituents attached to the Be atom, permitting us to modulate the electron-accepting ability of the BeXH or BeX2 molecule.
Figure
The association of pyrazole and imidazole with BeX2 derivatives dramatically enhances the acidity of the azole, so the complex imidazole:BeCl2 becomes a NH acid that is stronger than oxalic acid in the gas phase  相似文献   

18.
A relative complete study on the mechanisms of the proton transfer reactions of 2-thioxanthine was carried out with density functional theory. The models were designed with monohydrated and dihydrated microsolvent catalyses either with or without the presence of water solvent considered with the polarized continuum model (PCM). A total number of 114 complexes and 67 transition states were found with the B3LYP/6-311+G** calculations. The energies were refined with both B3LYP/aug-cc-pVTZ and PCM-B3LYP/aug-cc-pVTZ methods. The activation energies were reported with respect to the Gibbs free energies obtained in conjunction with the standard statistical thermodynamics. Possible reaction pathways were confirmed with the intrinsic reaction coordinates. Pathways via C8 atom on the imidazole ring, via the bridged C4 and C5 atoms between pyrimidine and imidazole rings and via N, O and S atom on the pyrimidine ring were examined. The results show that the most feasible pathway is the proton transfers within the long range solvent surrounding via the N, O and S atoms in the pyrimidine ring with di-hydrated catalysis: N(7)H?+?2H2O?→?IM89?→?IM90?→?P13?+?2H2O?→?IM91?→?IM92?→?P6?+?2H2O?→?IM71?→?IM72?→?P7?+?2H2O?→?IM107?→?IM108?→?P18?+?2H2O?→?IM111?→?IM112?→?P19?+?2H2O?→?IM113?→?IM114?→?P17?+?2H2O?→?IM105?→?IM106?→?N(9)H?+?2H2O that has the highest energy barrier of 44.0 kJ mol?1 in the transition of IM89 to IM90 via TS54. The small energy barrier is in good agreement with the experimental observation that 2-TX tautomerizes at room temperature in water. In the aqueous phase, the most stable intermediate is found to be IM21 [N(7)H?+?2H2O] and the possible co-existing species are the monohydrated IM1, IM9, IM39 and IM46, and the di-hydrated IM5, IM8, IM13, IM16, IM81, IM89, IM90, IM91 and IM106 complexes that have a relative concentration larger than 10?6 (1 ppm) with respect to IM21.
Figure
Mechanisms of the proton transfer reactions of 2-thioxanthine were investigated with both B3LYP/aug-cc-pVTZ//B3LYP/6-311+G** and PCM-B3LYP/aug-cc-pVTZ//B3LYP/6-311+G**. The models were designed with monohydrated and dihydrated microsolvent either with or without the presence of water solvent. The results show that the most feasible pathway is the reactions within the long range solvent surrounding via the N, O and S atoms in the pyrimidine ring with di-hydrated catalysis: N(7)H?+?2H2O?→?IM90?→?IM91?→?P13?+?2H2O?→?IM92?→?IM93?→?P6?+?2H2O?→?IM72?→?IM73?→?P7?+?2H2O?→?IM109?→?IM110?→?P18?+?2H2O?→?IM113?→?IM114?→?P19?+?2H2O?→?IM115?→?IM116?→?P17?+?2H2O?→?IM107?→?IM108?→?N(9)H?+?2H2O that has the highest barrier of 44.0 kJ mol?1 in the transition of IM90 to IM91 via TS54. The barrier is adequate for a reaction at room temperature that consists well with the experimental observations.  相似文献   

19.
To understand the chemical behavior of uranyl complexes in water, a bis-uranyl [(phen)(UO2)(μ2–F)(F)]2 (A; phen?=?phenanthroline, μ2?=?doubly bridged) and its hydrated form A?·?(H2O)n (n?=?2, 4 and 6) were examined using scalar relativistic density functional theory. The addition of water caused the phen ligands to deviate slightly from the U22–F)2 plane, and red-shifts the U–F-terminal and U?=?O stretching vibrations. Four types of hydrogen bonds are present in the optimized hydrated A?·?(H2O)n complexes; their energies were calculated to fall within the range 4.37–6.77 kcal mol-1, comparable to the typical values of 5.0 kcal mol-1 reported for hydrogen bonds. An aqueous environment simulated by explicit and/or implicit models lowers and re-arranges the orbitals of the bis-uranyl complex.
Figure
A bis(uranyl) complex [(phen)(UO2)(μ2–F)(F)]2 (A) and its solvated form A?·?(H2O)n were examined using scalar relativistic density functional theory. Hydrogen bonds cause the phen ligand to slightly deviate from the equatorial plane of the uranyl ion, resulting in a pronounced red-shift of the U–F-terminal and U?=?O asymmetric stretching vibrations. The calculated energies fall within 4.4?–6.8 kcal/mol. Explicit and/or implicit aqueous solvation re-arranges the molecular orbitals of the complex  相似文献   

20.
Density functional theory (DFT) with relativistic corrections of zero-order regular approximation (ZORA) has been applied to explore the reaction mechanisms of ethane dehydrogenation by Zr atom with triplet and singlet spin-states. Among the complicated minimum energy reaction path, the available states involves three transition states (TS), and four stationary states (1) to (4) and one intersystem crossing with spin-flip (marked by ?): 3 Zr + C 2 H 6 3 Zr-CH 3 -CH 3 ( 3 1) → 3 TS 1/2 3 ZrH-CH 2 -CH 3 ( 3 2) → 3 TS 2/3 ? 1 ZrH2-CH2 = CH2 ( 1 3) → 1 TS 3/4 1 ZrH 3 -CH = CH 2 ( 1 4). The minimum energy crossing point is determined with the help of the DFT fractional-occupation-number (FON) approach. The spin inversion leads the reaction pathway transferring from the triplet potential energy surface (PES) to the singlet’s accompanying with the activation of the second C-H bond. The overall reaction is calculated to be exothermic by about 231 kJ mol?1. Frequency and NBO analysis are also applied to confirm with the experimental observed data.
Reaction 3 Zr + C 2 H 6 → 3 ZrH ? CH 2 ? CH 3 ? 1 ZrH 2 ? CH 2 = CH 2 → 1 ZrH 3 ? CH = CH 2 $ {}^{\mathbf{3}}\mathrm{Zr}+{\mathrm{C}}_{\mathbf{2}}{\mathrm{H}}_{\mathbf{6}}{\to}^{\mathbf{3}}\mathrm{Zr}\mathrm{H}-{\mathrm{C}\mathrm{H}}_{\mathbf{2}}-{\mathrm{C}\mathrm{H}}_{\mathbf{3}}{\Rightarrow}^{\mathbf{1}}{\mathrm{ZrH}}_2-{\mathrm{C}\mathrm{H}}_2={\mathrm{C}\mathrm{H}}_2{\to}^{\mathbf{1}}{\mathrm{ZrH}}_{\mathbf{3}}-\mathrm{CH}={\mathrm{C}\mathrm{H}}_{\mathbf{2}} $ proceeds via spin-flip surface hopping over several transition states has been investigated. The minimum energy crossing point is determined with the help of the DFT fractional-occupation-number (FON) approach.  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