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 109 毫秒
1.
We report variational transition‐state theory calculations for the OH + O3→ HO2 + O2 reaction based on the recently reported double many‐body expansion potential energy surface for ground‐state HO4 [Chem Phys Lett 2000, 331, 474]. The barrier height of 1.884 kcal mol?1 is comparable to the value of 1.77–2.0 kcal mol?1 suggested by experimental measurements, both much smaller than the value of 2.16–5.11 kcal mol?1 predicted by previous ab initio calculations. The calculated rate constant shows good agreement with available experimental results and a previous theoretical dynamics prediction, thus implying that the previous ab initio calculations will significantly underestimate the rate constant. Variational and tunneling effects are found to be negligible over the temperature range 100–2000 K. The O1? O2 bond is shown to be spectator like during the reactive process, which confirms a previous theoretical dynamics prediction. © 2007 Wiley Periodicals, Inc. 39: 148–153, 2007  相似文献   

2.
The adsorption and the mechanism of the oxidative dehydrogenation (ODH) of propane over VO2‐exchanged MCM‐22 are investigated by DFT calculations using the M06‐L functional, which takes into account dispersion contributions to the energy. The adsorption energies of propane are in good agreement with those from computationally much more demanding MP2 calculations and with experimental results. In contrast, B3LYP binding energies are too small. The reaction begins with the movement of a methylene hydrogen atom to the oxygen atom of the VO2 group, which leads to an isopropyl radical bound to a HO? V? O intermediate. This step is rate determining with the apparent activation energy of 30.9 kcal mol?1, a value within the range of experimental results for ODH over other silica supports. In the propene formation step, the hydroxyl group is the more reactive group requiring an apparent activation energy of 27.7 kcal mol?1 compared to that of the oxy group of 40.8 kcal mol?1. To take the effect of the extended framework into account, single‐point calculations on 120T structures at the same level of theory are performed. The apparent activation energy is reduced to 28.5 kcal mol?1 by a stabilizing effect caused by the framework. Reoxidation of the catalyst is found to be important for the product release at the end of the reaction.  相似文献   

3.
In the title compound, C10H7NO3·H2O, the zwitterionic organic molecules and the water molecules are connected by N—H...O and O—H...O hydrogen bonds to form ribbons, and π–π stacking interactions expand these ribbons into a three‐dimensional net. The energies of these hydrogen bonds adopt values typical for mildly weak interactions (3.33–7.75 kcal mol−1; 1 kcal mol−1 = 4.184 kJ mol−1). The total π–π stacking interactions between aromatic molecules can be classified as mildly strong (energies of 15.3 and 33.9 kcal mol−1), and they are made up of multiple constituent π–π interactions between six‐membered rings. The short intermolecular C—H...O contact between two zwitterionic molecules is nonbonding in character.  相似文献   

4.
2‐Furaldehyde (2‐FA), also known as furfural or 2‐furancarboxaldehyde, is an heterocyclic aldehyde that can be obtained from the thermal dehydration of pentose monosaccharides. This molecule can be considered as an important sustainable intermediate for the preparation of a great variety of chemicals, pharmaceuticals and furan‐based polymers. Despite the great importance of this molecule, its gas‐phase basicity (GB) has never been measured. In this work, the GB of 2‐FA was determined by the extended Cooks's kinetic method from electrospray ionization triple quadrupole tandem mass spectrometric experiments along with theoretical calculations. As expected, computational results identify the aldehydic oxygen atom of 2‐FA as the preferred protonation site. The geometries of O‐O‐cis and O‐O‐trans 2‐FA and of their six different protomers were calculated at the B3LYP/aug‐TZV(d,p) level of theory; proton affinity (PA) values were also calculated at the G3(MP2, CCSD(T)) level of theory. The experimental PA was estimated to be 847.9 ± 3.8 kJ mol?1, the protonation entropy 115.1 ± 5.03 J mol?1 K?1 and the GB 813.6 ± 4.08 kJ mol?1 at 298 K. From the PA value, a ΔH°f of 533.0 ± 12.4 kJ mol?1 for protonated 2‐FA was derived. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

5.
The photochemistry of 1,2‐dihydro‐1,2‐azaborinine derivatives was studied under matrix isolation conditions and in solution. Photoisomerization occurs exclusively to the Dewar valence isomers upon irradiation with UV light (>280 nm) with high quantum yield (46 %). Further photolysis with UV light (254 nm) results in the formation of cyclobutadiene and an iminoborane derivative. The thermal electrocyclic ring‐opening reaction of the Dewar valence isomer back to the 1,2‐dihydro‐1‐tert‐butyldimethylsilyl‐2‐mesityl‐1,2‐azaborinine has an activation barrier of (27.0±1.2) kcal mol?1. In the presence of the Wilkinson catalyst, the ring opening occurs rapidly and exothermically (ΔH=(?48±1) kcal mol?1) at room temperature.  相似文献   

6.
CF3H as a proton donor was paired with a variety of anions, and its properties were assessed by MP2/aug‐cc‐pVDZ calculations. The binding energy of monoanions halide, NO3?, formate, acetate, HSO4?, and H2PO4? lie in the 12–17 kcal mol?1 range, although F? is more strongly bound, by 26 kcal mol?1. Dianions SO42? and HPO42? are bound by 27 kcal mol?1, and trianion PO43? by 45 kcal mol?1. When two O atoms are available on the anion, the CH???O? H‐bond (HB) is usually bifurcated, although asymmetrically. The CH bond is elongated and its stretching frequency redshifted in these ionic HBs, but the shift is reduced in the bifurcated structures. Slightly more than half of the binding energy is attributed to Coulombic attraction, with smaller contributions from induction and dispersion. The amount of charge transfer from the anions to the σ*(CH) orbital correlates with many of the other indicators of bond strength, such as binding energy, CH bond stretch, CH redshift, downfield NMR spectroscopic chemical shift of the bridging proton, and density at bond critical points.  相似文献   

7.
Treatment of the salt [PPh4]+[Cp*W(S)3]? ( 6 ) with allyl bromide gave the neutral complex [Cp*W(S)2S‐CH2‐CH?CH2] ( 7 ). The product 7 was characterized by an X‐ray crystal structure analysis. Complex 7 features dynamic NMR spectra that indicate a rapid allyl automerization process. From the analysis of the temperature‐dependent NMR spectra a Gibbs activation energy of ΔG (278 K)≈13.7±0.1 kcal mol?1 was obtained [ΔH≈10.4±0.1 kcal mol?1; ΔS≈?11.4 cal mol?1 K?1]. The DFT calculation identified an energetically unfavorable four‐membered transition state of the “forbidden” reaction and a favorable six‐membered transition state of the “Cope‐type” allyl rearrangement process at this transition‐metal complex core.  相似文献   

8.
A new preconcentration method is presented for lead on TAN‐loaded polyurethane foam (PUF) and its measurement by differential pulse anodic stripping voltammetry (DPASV). The optimum sorption conditions of 1.29 × 10?5 M solution of Pb(II) ions on TAN‐loaded PUF were investigated. The maximum sorption was observed at pH 7 with 20 minutes equilibrated time on 7.25 mg mL?1 of TAN‐loaded foam. The kinetic study indicates that the overall sorption process was controlled by the intra‐particle diffusion process. The validity of Freundlich, Langmuir and Dubinin ‐ Radushkevich adsorption isotherms were tested. The Freundlich constants 1/n and KF are evaluated to be 0.45 ±0.04 and (1.03 +0.61) × 10?3 mol g?1, respectively. The monolayer sorption capacity and adsorption constant related to the Langmuir isotherm are (1.38 ± 0.08) × 10?5 mol g?1 and (1.46 ± 0.27) × 105 L mol?1, respectively. The mean free energy of Pb(II) ions sorption on‐TAN loaded PUF is 11.04 ± 0.28 kJ mol?1 indicating chemisorption phenomena. The effect of temperature on the sorption yields thermodynamics parameters of ΔH, ΔS and ΔG at 298 K that are 15.0 ± 1.4 kJ mol?1, 74 ±5 J mol?1 K?1 and ‐7.37 ± 0.28 kJ mol?1, respectively. The positive values of enthalpy (ΔH) and entropy (ΔS) indicate the endothermic sorption and stability of the sorbed complexes are entropy driven. However, the negative value of Gibb's free energy (ΔG) indicates the spontaneous nature of sorption. On the basis of these data, the sorption mechanism has been postulated. The effect of different foreign ions on the sorption and desorption studies were also carried out. The method was successfully applied for the determination of lead from different water samples at ng levels.  相似文献   

9.
The potential energy surfaces of the naphthalene dimer and benzene–naphthalene complexes are investigated using the recently developed DFT/CCSD(T) correction scheme [J. Chem. Phys. 2008 , 128, 114 102]. One and three minima are located on the PES of the benzene–naphthalene and the naphthalene dimer complexes, respectively, all of which are of the parallel‐displaced type. The stabilities of benzene–naphthalene and the naphthalene dimer are ?4.2 and ?6.2 kcal mol?1, respectively. Unlike the benzene dimer, where the T‐shaped complex is the global minimum, the lowest‐energy T‐shaped structure is about 0.2 and 1.6 kcal mol?1 above the global minimum on the benzene–naphthalene and the naphthalene dimer potential energy surfaces, respectively.  相似文献   

10.
A common feature of several classes of intrinsically reactive proteins with diverse biological functions is that they undergo self‐catalyzed reactions initiated by an N→O or N→S acyl shift of a peptide bond adjacent to a serine, threonine, or cysteine residue. In this study, we examine the N→O acyl shift initiated peptide‐bond hydrolysis at the serine residue on a model compound, glycylserine (GlySer), by means of DFT and ab initio methods. In the most favorable rate‐determining transition state, the serine ?COO? group acts as a general base to accept a proton from the attacking ?OH function, which results in oxyoxazolidine ring closure. The calculated activation energy (29.4 kcal mol?1) is in excellent agreement with the experimental value, 29.4 kcal mol?1, determined by 1H NMR measurements. A reaction mechanism for the entire process of GlySer dipeptide hydrolysis is also proposed. In the case of proteins, we found that when no other groups that may act as a general base are available, the N→O acyl shift mechanism might instead involve a water‐assisted proton transfer from the attacking serine ?OH group to the amide oxygen. However, the calculated energy barrier for this process is relatively high (33.6 kcal mol?1), thus indicating that in absence of catalytic factors the peptide bond adjacent to serine is no longer a weak point in the protein backbone. An analogous rearrangement involving the amide N‐protonated form, rather than the principle zwitterion form of GlySer, was also considered as a model for the previously proposed mechanism of sea‐urchin sperm protein, enterokinase, and agrin (SEA) domain autoproteolysis. The calculated activation energy (14.3 kcal mol?1) is significantly lower than the experimental value reported for SEA (≈21 kcal mol?1), but is still in better agreement as compared to earlier theoretical attempts.  相似文献   

11.
The C?H activation in the tandem, “merry‐go‐round”, [(dppp)Rh]‐catalyzed (dppp=1,3‐bis(diphenylphosphino)propane), four‐fold addition of norborene to PhB(OH)2 has been postulated to occur by a C(alkyl)?H oxidative addition to square‐pyramidal RhIII?H species, which in turn undergoes a C(aryl)?H reductive elimination. Our DFT calculations confirm the RhI/RhIII mechanism. At the IEFPCM(toluene, 373.15 K)/PBE0/DGDZVP level of theory, the oxidative addition barrier was calculated to be 12.9 kcal mol?1, and that of reductive elimination was 5.0 kcal mol?1. The observed selectivity of the reaction correlates well with the relative energy barriers of the cycle steps. The higher barrier (20.9 kcal mol?1) for norbornyl–Rh protonation ensures that the reaction is steered towards the 1,4‐shift (total barrier of 16.3 kcal mol?1), acting as an equilibration shuttle. The carborhodation (13.2 kcal mol?1) proceeds through a lower barrier than the protonation (16.7 kcal mol?1) of the rearranged aryl–Rh species in the absence of o‐ or m‐substituents, ensuring multiple carborhodations take place. However, for 2,5‐dimethylphenyl, which was used as a model substrate, the barrier for carborhodation is increased to 19.4 kcal mol?1, explaining the observed termination of the reaction at 1,2,3,4‐tetra(exo‐norborn‐2‐yl)benzene. Finally, calculations with (Z)‐2‐butene gave a carborhodation barrier of 20.2 kcal mol?1, suggesting that carborhodation of non‐strained, open‐chain substrates would be disfavored relative to protonation.  相似文献   

12.
In this work, a density function theory (DFT) study is presented for the HNS/HSN isomerization assisted by 1–4 water molecules on the singlet state potential energy surface (PES). Two modes are considered to model the catalytic effect of these water molecules: (i) water molecule(s) participate directly in forming a proton transfer loop with HNS/HSN species, and (ii) water molecules are out of loop (referred to as out‐of‐loop waters) to assist the proton transfer. In the first mode, for the monohydration mechanism, the heat of reaction is 21.55 kcal · mol?1 at the B3LYP/6‐311++G** level. The corresponding forward/backward barrier lowerings are obtained as 24.41/24.32 kcal · mol?1 compared with the no‐water‐assisting isomerization barrier T (65.52/43.87 kcal · mol?1). But when adding one water molecule on the HNS, there is another special proton‐transfer isomerization pathway with a transition state 10T′ in which the water is out of the proton transfer loop. The corresponding forward/backward barriers are 65.89/65.89 kcal · mol?1. Clearly, this process is more difficult to follow than the R–T–P process. For the two‐water‐assisting mechanism, the heat of reaction is 19.61 kcal · mol?1, and the forward/backward barriers are 32.27/12.66 kcal · mol?1, decreased by 33.25/31.21 kcal · mol?1 compared with T. For trihydration and tetrahydration, the forward/backward barriers decrease as 32.00/12.60 (30T) and 37.38/17.26 (40T) kcal · mol?1, and the heat of reaction decreases by 19.39 and 19.23 kcal · mol?1, compared with T, respectively. But, when four water molecules are involved in the reactant loop, the corresponding energy aspects increase compared with those of the trihydration. The forward/backward barriers are increased by 5.38 and 4.66 kcal · mol?1 than the trihydration situation. In the second mode, the outer‐sphere water effect from the other water molecules directly H‐bonded to the loop is considered. When one to three water molecules attach to the looped water in one‐water in‐loop‐assisting proton transfer isomerization, their effects on the three energies are small, and the deviations are not more than 3 kcal · mol?1 compared with the original monohydration‐assisting case. When adding one or two water molecules on the dihydration‐assisting mechanism, and increasing one water molecule on the trihydration, the corresponding energies also are not obviously changed. The results indicate that the forward/backward barriers for the three in‐loop water‐assisting case are the lowest, and the surrounding water molecules (out‐of‐loop) yield only a small effect. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

13.
A new compound, 2‐(dinitromethylene)‐1,3‐diazacyclopentane (DNDZ), was prepared by the reaction of 1,1‐diamino‐2,2‐dinitroethylene (FOX‐7) with 1,2‐diaminoethane in N‐methylpyrrolidone (NMP). Thermal decomposition of DNDZ was studied under non‐isothermal conditions by DSC, TG/DTG methods, and the enthalpy, apparent activation energy and pre‐exponential factor of the exothermic decomposition reaction were obtained as 317.13 kJ·mol?1, 269.7 kJ·mol?1 and 1024.51 s?1, respectively. The critical temperature of thermal explosion was 261.04°C. Specific heat capacity of DNDZ was determined with a micro‐DSC method and a theoretical calculation method, and the molar heat capacity was 205.41 J·mol?1·K?1 at 298.15 K. Adiabatic time‐to‐explosion was calculated to be a certain value between 263–289 s. DNDZ has higher thermal stability than FOX‐7.  相似文献   

14.
Although supramolecular chemistry is traditionally an experimental discipline, computations have emerged as important tools for the understanding of supramolecules. We have explored how well commonly used density functional theory quantum mechanics and polarizable continuum solvation models can calculate binding affinities of host‐guest systems. We report the calculation of binding affinities for eight host–guest complexes and compare our results to experimentally measured binding free energies that span the range from ?2.3 to ?6.1 kcal mol?1. These systems consist of four hosts (biotin[6]uril, triphenoxymethane, cryptand, and bis‐thiourea) with different halide ions (F?, Cl?, Br?) in various media including organic and aqueous. The mean average deviation (MAD) of calculated from measured ΔGa is 2.5 kcal mol?1 when using B3LYP‐D3 with either CPCM or PCM. This MAD value lowers even more by eliminating two outliers: 1.1 kcal mol?1 for CPCM and 1.2 kcal mol?1 for PCM. The best DFT and implicit solvation model combination that we have studied is B3LYP?D3 with either CPCM or PCM.  相似文献   

15.
The gas‐phase elimination kinetics of the above‐mentioned compounds were determined in a static reaction system over the temperature range of 369–450.3°C and pressure range of 29–103.5 Torr. The reactions are homogeneous, unimolecular, and obey a first‐order rate law. The rate coefficients are given by the following Arrhenius expressions: ethyl 3‐(piperidin‐1‐yl) propionate, log k1(s?1) = (12.79 ± 0.16) ? (199.7 ± 2.0) kJ mol?1 (2.303 RT)?1; ethyl 1‐methylpiperidine‐3‐carboxylate, log k1(s?1) = (13.07 ± 0.12)–(212.8 ± 1.6) kJ mol?1 (2.303 RT)?1; ethyl piperidine‐3‐carboxylate, log k1(s?1) = (13.12 ± 0.13) ? (210.4 ± 1.7) kJ mol?1 (2.303 RT)?1; and 3‐piperidine carboxylic acid, log k1(s?1) = (14.24 ± 0.17) ? (234.4 ± 2.2) kJ mol?1 (2.303 RT)?1. The first step of decomposition of these esters is the formation of the corresponding carboxylic acids and ethylene through a concerted six‐membered cyclic transition state type of mechanism. The intermediate β‐amino acids decarboxylate as the α‐amino acids but in terms of a semipolar six‐membered cyclic transition state mechanism. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 106–114, 2006  相似文献   

16.
By using a set of model reactions, we estimated the heat of formation of gaseous UO22+ from quantum‐chemical reaction enthalpies and experimental heats of formation of reference species. For this purpose, we performed relativistic density functional calculations for the molecules UO22+, UO2, UF6, and UF5. We used two gradient‐corrected exchange‐correlation functionals (revised Perdew–Burke–Ernzerhof (PBEN) and Becke–Perdew (BP)) and we accounted for spin‐orbit interaction in a self‐consistent fashion. Indeed, spin‐orbit interaction notably affects the energies of the model reactions, especially if compounds of UIV are involved. Our resulting theoretical estimates for Δf (UO22+), 365±10 kcal mol?1 (PBEN) and 370±12 kcal mol?1 (BP), are in quantitative agreement with a recent experimental result, 364±15 kcal mol?1. Agreement between the results of the two different exchange‐correlation functionals PBEN and BP supports the reliability of our approach. The procedure applied offers a general means to derive unknown enthalpies of formation of actinide species based on the available well‐established data for other compounds of the element in question.  相似文献   

17.
A combined experimental and theoretical approach has been employed to establish the basicity and proton affinity of SiF4 and the structure of SiF4H+. The kinetics and energetics for the transfer of a proton between SiF4, N2, and Xe have been explored experimentally in helium at 0.35±0.02 torr and 297±3 K with a selected-ion flow tube apparatus. The results of equilibrium constant measurements are reported that provide a basicity and proton affinity for SiF4 at 297±3 K of 111.4±1.0 and 117.7±1.2 kcal mol?1, respectively. These values are more than 2.5 kcal mol?1 lower than currently recommended values. The basicity order was determined to be GB(Xe)>GB(SiF4)>GB(N2), while the proton-affinity order was shown to be PA(Xe)>PA(N2)>PA (SiF4). Ab initio molecular orbital computations at MP4SDTQ(fc)/6-311++G(3df,3pd) using geometries from B3LYP/6-31+G(d,p) indicate a value for PA(SiF4)=118.7 kcal mol?1 that is in good agreement with experiment. Also, the most stable structure of SiF4H+ is shown to correspond to a core SiF 3 + cation solvated by HF with a binding energy of 43. 9 kcal mol?1. Support for this structure is found in separate SIFT collision induced dissociation (CID) measurements that indicate exclusive loss of HF.  相似文献   

18.
A theoretical kinetic study of the thermal decomposition of 1‐chlorohexane in gas phase between 600 and 1000 K was performed. Transition‐state theory and unimolecular reaction rate theory were combined with molecular information provided by quantum chemical calculations. Particularly, the B3LYP, BMK, M05–2X, and M06–2X formulations of the density functional theory (DFT) and the high‐level ab initio methods G3B3 and G4 were employed. The possible reaction channels for the thermal decomposition of 1‐chlorohexane were investigated, and the reaction takes place through the elimination of HCl with the formation of 1‐hexene. The derived high‐pressure limit rate coefficients are k (600–1000 K) = (8 ± 5) × 1013 exp[‐((56.7 ± 0.4) kcal mol−1/RT )] s−1. The pressure effect over the reaction was analyzed from the calculation of the low‐pressure limit rate coefficients and the falloff curves. In addition, the standard enthalpies of formation at 298 K of −46.9 ± 1.5 kcal mol−1 for 1‐chlorohexane and 5.8 ± 1.5 kcal mol−1 for C6H13 radical were derived from isodesmic and isogiric reactions at high levels of theory.  相似文献   

19.
The thermal decomposition reaction of acetone cyclic diperoxide (3,3,6,6‐tetramethyl‐1,2,4,5‐tetroxane, ACDP), in the temperature range of 130.0–166.0°C and initial concentrations range of 0.4–3.1 × 10?2 mol kg?1 has been studied in methyl t‐butyl ether solution. The thermolysis follows first‐order kinetic laws up to at least ca 60% ACDP conversion. Under the experimental conditions, the activation parameters of the initial step of the reaction (ΔH# = 33.6 ± 1.1 kcal mol?1; ΔS# = ?4.1 ± 0.7 cal mol?1 K?1; ΔG# = 35.0 ± 1.1 kcal mol?1) and acetone, as the only organic product, support a stepwise reaction mechanism with the homolytic rupture of one of its peroxidic bond. Also, participation of solvent molecules in the reaction is postulated given an intermediate diradical, which further decomposes by C? O bond ruptures, yielding a stoichiometric amount of acetone (2 mol per mole of ACDP decomposed). The results are compared with those obtained for the above diperoxide thermolysis in other solvents. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 302–307, 2004  相似文献   

20.
Homobimetallic metallophilic interactions between copper, silver, and gold‐based [(NHC)MX]‐type complexes (NHC=N‐heterocyclic carbene, i.e, 1,3,4‐trimethyl‐4,5‐dihydro‐1H‐1,2,4‐triazol‐5‐ylidene; X=F, Cl, Br, I) were investigated by means of ab initio interaction energies, Ziegler–Rauk‐type energy‐decomposition analysis, the natural orbital for chemical valence (NOCV) framework, and the noncovalent interaction (NCI) index. It was found that the dimers of these complexes predominantly adopt a head‐to‐tail arrangement with typical M ??? M distance of 3.04–3.64 Å, in good agreement with the experimental X‐ray structure determined for [{(NHC)AuCl}2], which has an Au ??? Au distance of 3.33 Å. The interaction energies between silver‐ and gold‐based monomers are calculated to be about ?25 kcal mol?1, whereas that for the Cu congener is significantly lower (?19.7 kcal mol?1). With the inclusion of thermal and solvent contributions, both of which are destabilizing, by about 15 and 8 kcal mol?1, respectively, an equilibrium process is predicted for the formation of dimer complexes. Energy‐decomposition analysis revealed a dominant electrostatic contribution to the interaction energy, besides significantly stabilizing dispersion and orbital interactions. This electrostatic contribution is rationalized by NHC(δ+) ??? halogen(δ?) interactions between monomers, as demonstrated by electrostatic potentials and derived charges. The dominant NOCV orbital indicates weakening of the π backdonation in the monomers on dimer formation, whereas the second most dominant NOCV represents an electron‐density deformation according to the formation of a very weak M ??? M bond. One of the characteristic signals found in the reduced density gradient versus electron density diagram corresponds to the noncovalent interactions between the metal centers of the monomers in the NCI plots, which is the manifestation of metallophilic interaction.  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