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 reaction force and the electronic flux, first proposed by Toro-Labbé et al. (J Phys Chem A 103:4398, 1999) have been expressed by the existing conceptual DFT apparatus. The critical points (extremes) of the chemical potential, global hardness and softness have been identified by means of the existing and computable energy derivatives: the Hellman-Feynman force, nuclear reactivity and nuclear stiffness. Specific role of atoms at the reaction center has been unveiled by indicating an alternative method of calculation of the reaction force and the reaction electronic flux. The electron dipole polarizability on the IRC has been analyzed for the model reaction HF + CO→HCOF. The electron polarizability determined on the IRC α e (ξ) was found to be reasonably parallel to the global softness curve S(ξ). The softest state on the IRC (not TS) coincides with zero electronic flux.
Figure
Variation of the electronic dipole polarizability  相似文献   

2.
A synchronous, concerted chemical process is rigorously divided by the reaction force F(R), the negative gradient of V(R), into “reactant” and “product” regions which are dominated by structural changes and an intervening “transition” region which is electronically intensive. The reaction force constant κ(R), the second derivative of V(R), is negative throughout the transition region, not just at the nominal transition state, at which κ(R) has a minimum. This is consistent with experimental evidence that there is a transition region, not simply a specific point. We show graphically that significant nonsynchronicity in the process is associated with the development of a maximum of κ(R) in the transition region, which increases as the process becomes more nonsynchronous. (We speculate that for a nonconcerted process this maximum is actually positive.) Thus, κ(R) can serve as an indicator of the level of nonsynchronicity.
Figure
Profiles of potential energy V(R), reaction force F(R), and reaction force constant κ(R) along the intrinsic reaction coordinate R for a nonsynchronous concerted chemical reaction.  相似文献   

3.
The reaction force F(ξ) is the negative gradient of the potential energy of a chemical process along the intrinsic reaction coordinate ξ. We extend the rigorous concept of F(ξ) to the “activation strain model” of Bickelhaupt et al., to formulate the “strain” force F str(ξ) that retards a reaction and the “interaction” force F int(ξ) that drives it. These are investigated for a group of Diels-Alder cycloadditions. The results fully support the interpretation of the minimum of F(ξ) as defining the beginning of the transition from deformed reactants to eventual products.  相似文献   

4.
5.
Hydrolysis of nucleic acids is of fundamental importance in biological sciences. Kinetic and theoretical studies on different substrates wherein the phosphodiester bond combined with alkyl or aryl groups and sugar moiety have been the focus of attention in recent literature. The present work focuses on understanding the mechanism and energetics of alkali metal (Li, Na, and K) catalyzed hydrolysis of phosphodiester bond in modeled substrates including Thymidylyl (3′-O, 5′-S) thymidine phosphodiester (Tp-ST) (1), 3′-Thymidylyl (1-trifluoroethyl) phosphodiester (Tp-OCH2CF3) (2), 3′-Thymidylyl (o-cholorophenyl) phosphodiester (Tp-OPh(o-Cl)) (3) and 3′-Thymidylyl(p-nitrophenyl) phosphodiester (Tp-OPh(p-NO2)) (4) employing density functional theory. Theoretical calculations reveal that the reaction follows a single-step (ANDN) mechanism where nucleophile attack and leaving group departure take place simultaneously. Activation barrier for potassium catalyzed Tp-ST hydrolysis (12.0 kcal mol?1) has been nearly twice as large compared to that for hydrolysis incorporating lithium or sodium. Effect of solvent (water) on activation energies has further been analyzed by adding a water molecule to each metal ion of the substrate. It has been shown that activation barrier of phosphodiester hydrolysis correlates well with basicity of leaving group.
Figure
Phosphodiester bond in Tp‐ST (1), Tp‐OCH2CF3 (2) Tp‐OPh(o‐Cl) (3) and Tp‐OPh(p‐NO2) (4)  相似文献   

6.
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   

7.
The interactions of L-aminoglucosidic stereoisomers such as rhodostreptomycins A (Rho A) and B (Rho B) with cations (Mg2+, Ca2+, and H+) were studied by a quantum mechanical method that utilized DFT with B3LYP/6-311G**. Docking studies were also carried out in order to explore the surface recognition properties of L-aminoglucoside with respect to Mg2+ and Ca2+ ions under solvated and nonsolvated conditions. Although both of the stereoisomers possess similar physicochemical/antibiotic properties against Helicobacter pylori, the thermochemical values for these complexes showed that its high affinity for Mg2+ cations caused the hydration of Rho B. According to the results of the calculations, for Rho A–Ca2+(H2O)6, ΔH = ?72.21 kcal?mol?1; for Rho B–Ca2+(H2O)6, ΔH = ?72.53 kcal?mol?1; for Rho A–Mg2+(H2O)6, ΔH = ?72.99  kcal?mol?1 and for Rho B–Mg2+(H2O)6, ΔH = ?95.00  kcal?mol?1, confirming that Rho B binds most strongly with hydrated Mg2+, considering the energy associated with this binding process. This result suggests that Rho B forms a more stable complex than its isomer does with magnesium ion. Docking results show that both of these rhodostreptomycin molecules bind to solvated Ca2+ or Mg2+ through hydrogen bonding. Finally, Rho B is more stable than Rho A when protonation occurs.
Figure
Rho B–H showed higher stability since it is considered a proton pump inhibitor, and is therefore a stronger inhibitor of Helicobacter pylori  相似文献   

8.
Theoretical investigations were carried out on the multi-channel reactions CF3 + SiHF3, CF3 + SiHCl3, CH3 + SiHF3, and CH3 + SiHCl3. Electronic structures were calculated at the MP2/6-311+G(d,p) level, and energetic information further refined by the MC-QCISD (single-point) method. The rate constants for major reaction channels were calculated by the canonical variational transition state theory with small-curvature tunneling correction over the temperature range of 200–1,500 K. The theoretical rate constants were in good agreement with the available experimental data and were fitted to the three parameter expression: k 1a(T) = 2.93 × 10?26 T 4.25 exp (?318.68/T), and k 2a(T) = 3.67 × 10?22 T 2.72 exp (?1,414.22/T), k 3a (T) = 7.00 × 10?24 T 3.27 exp (?384.04/T), k 4a(T) = 6.35 × 10?22 T 2.59 exp (?603.18/T) (in unit of cm3molecule?1s?1) are given. Our calculations indicate that hydrogen abstraction channel is the major channel due to the smaller barrier height among four channels considered.
Figure
Theoretical investigations on the reaction mechanisms of SiHX3 with CF3 and CH3 radicals. Rate constants were calculated in the temperature range 200―1,500 K. Our calculations indicate that hydrogen abstraction is the major channel, and is important in a wide variety of materials synthesis processes, in glow discharge deposition of amorphous silicon films, and in the semiconductor manufacturing process  相似文献   

9.
10.
Density functional theory (DFT) was used to investigate the ruthenium hydride-catalyzed regioselective addition reactions of benzaldehyde to isoprene leading to the branched β,γ-unsaturated ketone. All intermediates and the transition states were optimized completely at the B3LYP/6-31?G(d,p) level (LANL2DZ(f) for Ru, LANL2DZ(d) for P and Cl). Calculated results indicated that three catalysts RuHCl(CO)(PMe3)3 (1), RuH2(CO)(PMe3)3 (2), and RuHCl(PMe3)3 (3) exhibited different catalysis, and the first was the most excellent. The most favorable reaction pathway included the coordination of 1 to the less substituted olefin of isoprene, a hydrogen transfer reaction from ruthenium to the carbon atom C1, the complexation of benzaldehyde to ruthenium, the carbonyl addition, and the hydride elimination reaction. The carbonyl addition was the rate-determining step. The dominant product was the branched β,γ-unsaturated ketone. Furthermore, the presence of one toluene molecule lowered the activation free energy of the transition state of the carbonyl addition by hydrogen bonds between the protons of toluene and the chlorine, carbonyl oxygen of the ruthenium complex. On the whole, the solvent effect decreased the free energies of the species.
Figure
DFT study suggests that RuHCl(CO)(PMe3)3 has better catalysis than RuH2(CO)(PMe3)3 and RuHCl(PMe3)3 in the regioselective addition reactions of benzaldehyde to isoprene leading to the branched β,γ-unsaturated ketone.  相似文献   

11.
12.
13.
In this study, we performed several DFT, MP2, and BD(T) calculations on the 1,2-H shift reactions of two diaminocarbenes (1, 2) and a diamidocarbene (3) using the Gaussian 09 program. In Gaussian 09, the BD(T) method keyword requests a Brueckner doubles calculation including a perturbative triples contribution. Although N-heterocyclic carbenes (NHC) are typically known for their exceptional σ-donor abilities, recent studies have indicated that π-interactions also play a role in the bonding between NHCs and transition metals or BX3 (X = H, OH, NH2, CH3, CN, NC, F, Cl, and Br) (Nemcsok et al. Organomet 23:3640–3646, 2004, Esrafili. J Mol Model 18:2003–2011, 2012). In order to study the importance of π-interactions between carbenes and transition metals, Hobbs and co-workers (Hobbs et al. New J Chem 34:1295–1308, 2010) focused on the synthesis of NHCs with reduced-energy lowest unoccupied molecular orbitals. By introducing an oxalamide moiety into the heterocyclic backbone, they found the resulting carbene possessed higher electrophilicity than usual NHCs. According to our results, the N,N'-diamidocarbene should be more stable than the diaminocarbenes with respect to the 1,2-H shift reaction.
In this study, we performed several DFT, MP2, and BD(T) calculations on the 1,2-H shift reactions of two diaminocarbenes (1, 2) and a diamidocarbene (3). According to our results, the N,N'-diamidocarbene should be more stable than the diaminocarbenes with respect to the 1,2-H shift reaction. Due to the synthetic utility of N,N′-diamidocarbenes, we believe that our results could provide information to better rationalize their reactivity.  相似文献   

14.
A stochastic exploration of the quantum conformational spaces in the microsolvation of divalent cations with explicit consideration of up to six solvent molecules [Mg (H 2 O) n )]2+, (n?=?3, 4, 5, 6) at the B3LYP, MP2, CCSD(T) levels is presented. We find several cases in which the formal charge in Mg2+ causes dissociation of water molecules in the first solvation shell, leaving a hydroxide ion available to interact with the central cation, the released proton being transferred to outer solvation shells in a Grotthus type mechanism; this particular finding sheds light on the capacity of Mg2+ to promote formation of hydroxide anions, a process necessary to regulate proton transfer in enzymes with exonuclease activity. Two distinct types of hydrogen bonds, scattered over a wide range of distances (1.35–2.15 Å) were identified. We find that in inner solvation shells, where hydrogen bond networks are severely disturbed, most of the interaction energies come from electrostatic and polarization+charge transfer, while in outer solvation shells the situation approximates that of pure water clusters.
Figure
Water dissociation in the first solvation shell is observed only for [Mg(H2O)n]2+ clusters. The dissociated proton is then transferred to higher solvation shells via a Grotthus type mechanism  相似文献   

15.
The possibility of a new endohedral fullerene with a trapped aluminum carbide cluster, Al4C @C80-I h , was theoretical investigated. The geometries and electronic properties of it were investigated using density functional theory methods. The Al4C unit formally transfers six electrons to the C80 cage which induces stabilization of Al4C@C80. A favorable binding energy, relatively large HOMO-LUMO gap, electron affinities and ionization potentials suggested the Al4C@C80 is rather stable. The analysis of vertical ionization potential and vertical electron affinity indicate Al4C@C80 is a good electron acceptor.
Figure
An endohedral fullerene with a trapped aluminum carbide cluster, Al4C @C80-I h , was investigated using density functional theory. A favorable binding energy, relatively large HOMO-LUMO gap, electron affinities and ionization potentials suggested it is rather stable  相似文献   

16.
Calculations performed at the ab initio level using the recently reported planar concentric π-aromatic B18H6 2+(1) [Chen Q et al. (2011) Phys Chem Chem Phys 13:20620] as a building block suggest the possible existence of a new class of B3n H m polycyclic aromatic hydroboron (PAHB) clusters—B30H8(2), B39H9 2?(3), B42H10(4/5), B48H10(6), and B72H12(7)—which appear to be the inorganic analogs of the corresponding C n H m polycyclic aromatic hydrocarbon (PAHC) molecules naphthalene C10H8, phenalenyl anion C13H9 ?, phenanthrene/anthracene C14H10, pyrene C16H10, and coronene C24H12, respectively, in a universal atomic ratio of B:C?=?3:1. Detailed canonical molecular orbital (CMO), adaptive natural density partitioning (AdNDP), and electron localization function (ELF) analyses indicate that, as they are hydrogenated fragments of a boron snub sheet [Zope RR, Baruah T (2010) Chem Phys Lett 501:193], these PAHB clusters are aromatic in nature, and exhibit the formation of islands of both σ- and π-aromaticity. The predicted ionization potentials of PAHB neutrals and electron detachment energies of small PAHB monoanions should permit them to be characterized experimentally in the future. The results obtained in this work expand the domain of planar boron-based clusters to a region well beyond B20, and experimental syntheses of these snub B3n H m clusters through partial hydrogenation of the corresponding bare B3n may open up a new area of boron chemistry parallel to that of PAHCs in carbon chemistry.
Figure
Ab initio calculations predict the existence of polycyclic aromatic hydroboron clusters as fragments of a boron snub sheet; these clusters are analogs of polycyclic aromatic hydrocarbons  相似文献   

17.
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   

18.
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   

19.
Gas-phase reactions of ClO/BrO with RCl (R = CH3, C2H5, and C3H7) have been investigated in detail using the popular DFT functional BHandHLYP/aug-cc-pVDZ level of theory. As a result, our findings strongly suggest that the type of reaction is firstly initiated by a typical SN2 fashion. Subsequently, two competitive substitution steps, named as SN2-induced substitution and SN2-induced elimination, respectively, would proceed before the initial SN2 product ion-dipole complex separates, in which the former exhibits less reactivity than the latter. Those are consistent with relevant experimental results. Moreover, we have also explored reactivity difference for the title reactions in term of some factors derived from methyl group, p-π electronic conjugation, ionization energy (IE), as well as molecular orbital (MO) analysis.
Figure
Energy profiles for the ClO– reactions and BrO–reactions, respectively  相似文献   

20.
We report a DFT, TDDFT and DFTB investigation of the performance of two donor-π-acceptor (D-π-A)-type organic dyes bearing different electron-withdrawing groups (EWG) for dye-sensitized solar cells (DSSCs) to evaluate which EWG is better for an acrylic acid acceptor, i.e., Cyano (–CN) or o-nitrophenyl (o-NO2–Ph). A series of theoretical criteria applied successfully in our previous work to explain the different performance of organic dyes related to open-circuit photovoltage (V oc) and short-circuit current density (J sc) were used to evaluate the performance of the dyes with just different EWG. Our calculated results reveal that dye 2 with o-NO2–Ph has a larger vertical dipole moment, more electrons transferred from the dye to the semiconductor and a lower degree of charge recombination, which could lead to larger V oc; while the larger driving force and comparable light harvesting efficiency could lead to higher J sc , highlighting the potential of o-NO2–Ph as an EWG in an acrylic acid acceptor.
Figure
CN or o-NO2-Ph? Which is better for acrylic acid acceptor of donor-π-acceptor (D-π-A) dyes used in dye-sensitized solar cells (DSSCs) has been evaluated by DFT/TDDFT calculations.  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