首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Enthalpy, activation energy, and rate constant of 9 alkyl, 3 acyl, 3 alkoxyl, and 9 peroxyl radicals with alkanethiols, benzenethiol, and L ‐cysteine are calculated. The intersection parabolas model is used for activation energy calculations. Depending on the structure of attacking radical, the activation energy of reactions with alkylthiols varies from 3 to 43 kJ mol?1 for alkyl radicals, from 7 to 9 kJ mol?1 for alkoxyl, and from 18 to 35 kJ mol?1 for peroxyl radicals. The influence of adjacent π‐bonds on activation energy is estimated. The polar effect is found in reactions of hydroxyalkyl and acyl radicals with alkylthiols. The steric effect is observed in reactions of alkyl radicals with tert‐alkylthiols. All these factors are characterized via increments of activation energy. Quantum chemical calculations of activation energy and geometry of transition state were performed for model reactions: C?H3 + CH3SH, CH3O? + CH3SH, and HO2? + CH3SH with using density functional theory and Gaussian‐98. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 284–293, 2009  相似文献   

2.
The mechanisms for the reaction of CH3SSCH3 with OH radical are investigated at the QCISD(T)/6‐311++G(d,p)//B3LYP/6‐311++G(d,p) level of theory. Five channels have been obtained and six transition state structures have been located for the title reaction. The initial association between CH3SSCH3 and OH, which forms two low‐energy adducts named as CH3S(OH)SCH3 (IM1 and IM2), is confirmed to be a barrierless process, The S? S bond rupture and H? S bond formation of IM1 lead to the products P1(CH3SH + CH3SO) with a barrier height of 40.00 kJ mol?1. The reaction energy of Path 1 is ?74.04 kJ mol?1. P1 is the most abundant in view of both thermodynamics and dynamics. In addition, IMs can lead to the products P2 (CH3S + CH3SOH), P3 (H2O + CH2S + CH3S), P4 (CH3 + CH3SSOH), and P5 (CH4 + CH3SSO) by addition‐elimination or hydrogen abstraction mechanism. All products are thermodynamically favorable except for P4 (CH3 + CH3SSOH). The reaction energies of Path 2, Path 3, Path 4, and Path 5 are ?28.42, ?46.90, 28.03, and ?89.47 kJ mol?1, respectively. Path 5 is the least favorable channel despite its largest exothermicity (?89.47 kJ mol?1) because this process must undergo two barriers of TS5 (109.0 kJ mol?1) and TS6 (25.49 kJ mol?1). Hopefully, the results presented in this study may provide helpful information on deep insight into the reaction mechanism. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

3.
Ab initio molecular orbital calculations with moderately large polarization basis sets and including valence-electron correlation have been used to examine the structure and dissociation mechanisms of protonated methanol [CH3OH2]+. Stable isomers and transition structures have been characterized using gradient techniques. Protonated methanol is found to be the only stable isomer in the [CH5O]+ potential surface. There is no evidence for a tightly-bound complex, [HOCH2]+…?H2, analogous to the preferred structure [CH3]+…?H2 of [CH5]+. Protonated methanol is found to possess a pyramidal arrangement of bonds at the oxygen atom with a barrier to inversion of 8kJ mol?1. The lowest energy fragmentation pathways are dissociation into methyl cation and water (predicted to require 284 kJ mol?1 with zero reverse activation energy) and loss of molecular hydrogen (endothermic by 138 kJ mol?1 but with a reverse activation barrier of 149 kJ mol?1). The results offer a possible explanation as to why production of [CH2OH]+ from the reaction of methyl cation with water is not observed. Other dissociation processes examined include loss of a hydrogen atom to yield the methylenoxonium radical cation or methanol radical cation (requiring 441 and 490 kJ mol?1, respectively) and loss of a proton to yield neutral methanol (requiring 784 kJ mol?1).  相似文献   

4.
[RuCl2(NCCH3)2(cod)], an alternative starting material to [RuCl2(cod)] n for the preparation of ruthenium(II) complexes, has been prepared from the polymer compound and isolated in yields up to 87% using a new work-up procedure. The compound has been obtained as a yellow solid without water of crystallization. The complexes [RuCl2(NCR)2(cod)] spontaneously transform into dimers [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph). 1H NMR kinetic experiments for these transformations evidenced first-order behavior. [RuCl2(NCPh)2(cod)] dimerizes slower by a factor of ten than [RuCl2(NCCH3)2(cod)]. The following activation parameters, ΔH #?=?114?±?3?kJ?mol?1 and ΔS #?=?66?±?9?J?K?1?mol?1 for R?=?CH3CN (ΔG #?=?94?±?5?kJ?mol?1, 298.15?K) and ΔH #?=?122?±?2?kJ?mol?1 and ΔS #?=?75?±?6?J?K?1?mol?1 for R?=?Ph (ΔG #?=?100?±?4?kJ?mol?1, 298.15?K), have been calculated from the first-order rate constants in the temperature range 294–323?K. The kinetic parameters are in agreement with a two-step mechanism with dissociation of acetonitrile as the rate-determining step. The molecular structures of [Ru2Cl(μ-Cl)3(cod)2(NCR)] (R?=?Me, Ph) have been determined by X-ray diffraction.  相似文献   

5.
Complete dehydrogenation of methane is studied on model Pt catalysts by means of state‐of‐the‐art DFT methods and by a combination of supersonic molecular beams with high‐resolution photoelectron spectroscopy. The DFT results predict that intermediate species like CH3 and CH2 are specially stabilized at sites located at particles edges and corners by an amount of 50–80 kJ mol?1. This stabilization is caused by an enhanced activity of low‐coordinated sites accompanied by their special flexibility to accommodate adsorbates. The kinetics of the complete dehydrogenation of methane is substantially modified according to the reaction energy profiles when switching from Pt(111) extended surfaces to Pt nanoparticles. The CH3 and CH2 formation steps are endothermic on Pt(111) but markedly exothermic on Pt79. An important decrease of the reaction barriers is observed in the latter case with values of approximately 60 kJ mol?1 for first C? H bond scission and 40 kJ mol?1 for methyl decomposition. DFT predictions are experimentally confirmed by methane decomposition on Pt nanoparticles supported on an ordered CeO2 film on Cu(111). It is shown that CH3 generated on the Pt nanoparticles undergoes spontaneous dehydrogenation at 100 K. This is in sharp contrast to previous results on Pt single‐crystal surfaces in which CH3 was stable up to much higher temperatures. This result underlines the critical role of particle edge sites in methane activation and dehydrogenation.  相似文献   

6.
The potential energy surface for the [CH5N] system has been investigated using ab initio molecular orbital calculations with large, polarization basis sets and incorporating valence-electron correlation. Two [CH5N] isomers can be distinguished: the well known methylamine radical cation, [CH3NH2], and the less familiar methylenammonium radical cation, [CH2NH3]. The latter is calculated to lie 8 kJ mol?1 lower in energy. A substantial barrier (176 kJ mol?1) is predicted for rearrangement of [CH2NH3] to [CH3NH2]. In addition, a large barrier (202 kJ mol?1) is found for loss of a hydrogen radical from [CH2NH3] via direct N—H bond cleavage to give the aminomethyl cation [CH2NH2]+. These results are consistent with the existence of the methylenammonium ion [CH2NH3] as a stable observable species. The barrier to loss of a hydrogen radical from [CH3NH2] is calculated to be 140 kJ mol?1.  相似文献   

7.
Cation‐radicals and dications corresponding to hydrogen atom adducts to N‐terminus‐protonated Nα‐glycylphenylalanine amide (Gly‐Phe‐NH2) are studied by combined density functional theory and Møller‐Plesset perturbational computations (B3‐MP2) as models for electron‐capture dissociation of peptide bonds and elimination of side‐chain groups in gas‐phase peptide ions. Several structures are identified as local energy minima including isomeric aminoketyl cation‐radicals, and hydrogen‐bonded ion‐radicals, and ylid‐cation‐radical complexes. The hydrogen‐bonded complexes are substantially more stable than the classical aminoketyl structures. Dissociations of the peptide N? Cα bonds in aminoketyl cation‐radicals are 18–47 kJ mol?1 exothermic and require low activation energies to produce ion‐radical complexes as stable intermediates. Loss of the side‐chain benzyl group is calculated to be 44 kJ mol?1 endothermic and requires 68 kJ mol?1 activation energy. Rice‐Ramsperger‐Kassel‐Marcus (RRKM) and transition‐state theory (TST) calculations of unimolecular rate constants predict fast preferential N? Cα bond cleavage resulting in isomerization to ion‐molecule complexes, while dissociation of the Cα? CH2C6H5 bond is much slower. Because of the very low activation energies, the peptide bond dissociations are predicted to be fast in peptide cation‐radicals that have thermal (298 K) energies and thus behave ergodically. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

8.
This paper reports the computed O2 binding to heme, which for the first time explains experimental enthalpies for this process of central importance to bioinorganic chemistry. All four spin states along the relaxed Fe? O2‐binding curves were optimized using the full heme system with dispersion, thermodynamic, and scalar‐relativistic corrections, applying several density functionals. When including all these physical terms, the experimental enthalpy of O2 binding (?59 kJ mol?1) is closely reproduced by TPSSh‐D3 (?66 kJ mol?1). Dispersion changes the potential energy surfaces and leads to the correct electronic singlet and heptet states for bound and dissociated O2. The experimental activation enthalpy of dissociation (~82 kJ mol?1) was also accurately computed (~75 kJ mol?1) with an actual barrier height of ~60 kJ mol?1 plus a vibrational component of ~10 and ~5 kJ mol?1 due to the spin‐forbidden nature of the process, explaining the experimentally observed difference of ~20 kJ mol?1 in enthalpies of binding and activation. Most importantly, the work shows how the nearly degenerate singlet and triplet states increase crossover probability up to ~0.5 and accelerate binding by ~100 times, explaining why the spin‐forbidden binding of O2 to heme, so fundamental to higher life forms, is fast and reversible.  相似文献   

9.
A series of unusual chemical‐bond transformations were observed in the reactions of high active yttrium? dialkyl complexes with unsaturated small molecules. The reaction of scorpionate‐anchored yttrium? dibenzyl complex [TpMe2Y(CH2Ph)2(thf)] ( 1 , TpMe2=tri(3,5‐dimethylpyrazolyl)borate) with phenyl isothiocyanate led to C?S bond cleavage to give a cubane‐type yttrium–sulfur cluster, {TpMe2Y(μ3‐S)}4 ( 2 ), accompanied by the elimination of PhN?C(CH2Ph)2. However, compound 1 reacted with phenyl isocyanate to afford a C(sp3)? H activation product, [TpMe2Y(thf){μ‐η13‐OC(CHPh)NPh}{μ‐η32‐OC(CHPh)NPh}YTpMe2] ( 3 ). Moreover, compound 1 reacted with phenylacetonitrile at room temperature to produce γ‐deprotonation product [(TpMe2)2Y]+[TpMe2Y(N=C?CHPh)3]? ( 6 ), in which the newly formed N?C?CHPh ligands bound to the metal through the terminal nitrogen atoms. When this reaction was carried out in toluene at 120 °C, it gave a tandem γ‐deprotonation/insertion/partial‐TpMe2‐degradation product, [(TpMe2Y)2(μ‐Pz)2{μ‐η13‐NC(CH2Ph)CHPh}] ( 7 , Pz=3,5‐dimethylpyrazolyl).  相似文献   

10.
Ab initio molecular orbital calculations have been used to study the condensation reactions of CH3? with NH3, H2O, HF and H2S. Geometry optimization has been carried out at the Hartree—Fock (HF) level with the split-valence plus d-polarization 6-31G* basis set and improved relative energies obtained from calculations which employ the split-valence plus dp-polarization 6-31G** basis set with electron correlation incorporated via Moller—Plesset perturbation theory terminated at third order (MP3). Zero-point vibrational energies have also been determined and taken into account in deriving relative energies. The structures of the intermediates CH3XH? (X = NH2, OH, F and SH) have been obtained and dissociation of these intermediates into CH2X+ + H2 on the one hand, and CH3? + HX on the other, has been examined. It is found that for those species for which the methyl condensation reaction is observed to have an appreciable rate (X = NH2 and SH), the transition structure for hydrogen elimination from CH3XH? lies significantly lower in energy than the reactants CH3? + HX (by 75 and 70 kJ mol?1 respectively). On the other hand, for those species for which the methyl condensation reaction is not observed (X = OH and F), the transition structure for H2 elimination lies higher in energy than CH3? + HX (by 6 and 87 kJ mol?1 respectively).  相似文献   

11.
Density functional calculations yield energy barriers for H abstraction by oxygen radical sites in Li‐doped MgO that are much smaller (12±6 kJ mol?1) than the barriers inferred from different experimental studies (80–160 kJ mol?1). This raises further doubts that the Li+O.? site is the active site as postulated by Lunsford. From temperature‐programmed oxidative coupling reactions of methane (OCM), we conclude that the same sites are responsible for the activation of CH4 on both Li‐doped MgO and pure MgO catalysts. For a MgO catalyst prepared by sol–gel synthesis, the activity proved to be very different in the initial phase of the OCM reaction and in the steady state. This was accompanied by substantial morphological changes and restructuring of the terminations as transmission electron microscopy revealed. Further calculations on cluster models showed that CH4 binds heterolytically on Mg2+O2? sites at steps and corners, and that the homolytic release of methyl radicals into the gas phase will happen only in the presence of O2.  相似文献   

12.
The spontaneous self‐assembly of a neutral circular trinuclear TiIV‐based helicate is described through the reaction of titanium(IV) isopropoxide with a rationally designed tetraphenolic ligand. The trimeric ring helicate was obtained after diffusion of n‐pentane into a solution with dichloromethane. The circular helicate has been characterized by using single‐crystal X‐ray diffraction study, 13C CP‐MAS NMR and 1H NMR DOSY solution spectroscopic, and positive electrospray ionization mass‐spectrometric analysis. These analytical data were compared with those obtained from a previously reported double‐stranded helicate that crystallizes in toluene. The trimeric ring was unstable in a pure solution with dichloromethane and transformed into the double‐stranded helicate. Thermodynamic analysis by means of the PACHA software revealed that formation of the double‐stranded helicates was characterized by ΔH(toluene)=?30 kJ mol?1 and ΔS(toluene)=+357 J K?1 mol?1, whereas these values were ΔH(CH2Cl2)=?75 kJ mol?1 and ΔS(CH2Cl2)=?37 J K?1 mol?1 for the ring helicate. The transformation of the ring helicate into the double‐stranded helicate was a strongly endothermic process characterized by ΔH(CH2Cl2)=+127 kJ mol?1 and ΔH(n‐pentane)=+644 kJ mol?1 associated with a large positive entropy change ΔS=+1115 J K?1?mol?1. Consequently, the instability of the ring helicate in pure dichloromethane was attributed to the rather high dielectric constant and dipole moment of dichloromethane relative to n‐pentane. Suggestions for increasing the stability of the ring helicate are given.  相似文献   

13.
A three‐dimensional (3D) cage‐like organic network (3D‐CON) structure synthesized by the straightforward condensation of building blocks designed with gas adsorption properties is presented. The 3D‐CON can be prepared using an easy but powerful route, which is essential for commercial scale‐up. The resulting fused aromatic 3D‐CON exhibited a high Brunauer–Emmett–Teller (BET) specific surface area of up to 2247 m2 g?1. More importantly, the 3D‐CON displayed outstanding low pressure hydrogen (H2, 2.64 wt %, 1.0 bar and 77 K), methane (CH4, 2.4 wt %, 1.0 bar and 273 K), and carbon dioxide (CO2, 26.7 wt %, 1.0 bar and 273 K) uptake with a high isosteric heat of adsorption (H2, 8.10 kJ mol?1; CH4, 18.72 kJ mol?1; CO2, 31.87 kJ mol?1). These values are among the best reported for organic networks with high thermal stability (ca. 600 °C).  相似文献   

14.
A series of new indanimine ligands [ArN?CC2H3(CH3)C6H2(R)OH] (Ar = Ph, R = Me ( 1 ), R = H ( 2 ), and R = Cl ( 3 ); Ar = 2,6‐i‐Pr2C6H3, R = Me ( 4 ), R = H ( 5 ), and R = Cl ( 6 )) were synthesized and characterized. Reaction of indanimines with Ni(OAc)2·4H2O results in the formation of the trinuclear hexa(indaniminato)tri (nickel(II)) complexes Ni3[ArN = CC2H3(CH3)C6H2(R)O]6 (Ar = Ph, R = Me ( 7 ), R = H ( 8 ), and R = Cl ( 9 )) and the mononuclear bis(indaniminato)nickel (II) complexes Ni[ArN?CC2H3(CH3)C6H2(R)O]2 (Ar = 2,6‐i‐Pr2C6H3, R = Me ( 10 ), R = H ( 11 ), and R = Cl ( 12 )). All nickel complexes were characterized by their IR, NMR spectra, and elemental analyses. In addition, X‐ray structure analyses were performed for complexes 7 , 10 , 11 , and 12 . After being activated with methylaluminoxane (MAO), these nickel(II) complexes can polymerize norbornene to produce addition‐type polynorbornene (PNB) with high molecular weight Mv (106 g mol?1), highly catalytic activities up to 2.18 × 107 gPNB mol?1 Ni h?1. Catalytic activities and the molecular weight of PNB have been investigated for various reaction conditions. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 489–500, 2008  相似文献   

15.
Ab initio Molecular orbital calculations with large basis sets and incorporating correlation are used to examine the structures and relative energies of the vinyloxonium (CH2CHOH2+) and 1-hydroxyethyl (CH3CHOH+) cations. The best structure of the vinyloxonium ion has the OH2 plane perpendicular to the CCO plane. The energy difference between the vinyloxonium and 1-hydroxyethyl cations is predicted to be 92 kJ mol?1, substantially greater than a recent experimental estimate of 41 ± 12 kJ mol?1  相似文献   

16.
Mechanistic questions regarding the reductive cleavage of sulfonium ions by the NiI form of coenzyme F430 pentamethyl ester (F430M) were addressed in a series of kinetic studies and isotope labeling experiments. In neat DMF, methane formation from dialkyl(methyl)sulfonium ions consistently showed a delay time of ca. 1 h. In the presence of excess propanethiol, no delay was observed and methane formation followed pseudo-first-order kinetics with a logarithmic dependence of the initial rate on the concentration of propanethiol. From the temperature dependence of the reaction rate, an estimate for the activation parameters of ΔH# = 49 kJ mol?1 and (apparent) ΔS# = –114 J K?1 mol?1 was derived. The observation of deuterium incorporation into methane from (CH3)2CHOD, but not from (CH3)2CDOH, indicates that the fourth H-entity is introduced into CH4 as a proton, and that free CH3 radicals are not involved. In contrast to the reaction with the homogeneous one-electron reductant sodium naphthalide, the F430M-catalyzed reduction of mixed dialkyl(methyl)sulfonium ions showed a pronounced selectivity for the cleavage of Me? S over that of alkyl-S (alkyl ≠ Me) bonds. Mechanisms that are consistent with these results, as well as possible explanations for the time delay and the apparent highly negative entropy of activation, are discussed.  相似文献   

17.
18.
The volatilisation of ferrocene (Fc), dissolved in the ionic liquid N‐butyl‐N‐methylpyrrolidinium bis(trifluoromethylsulfonyl)imide, [C4mpyrr][NTf2], to the gas phase has been indirectly monitored by cyclic voltammetry and chronoamperometry. Simulation of the observed trends in concentration with time using a simple model allowed quantification of the process. Volatilisation of dissolved Fc under flowing wet and dry dinitrogen gas (N2) was found to be kinetically limited with a rate constant in the region of 2×10?7 cm s?1. The activation energy of diffusion for Fc was found to be 28.2±0.7 kJ mol?1, while the activation energy of volatilisation of Fc from [C4mpyrr][NTf2] to dry N2 was found to be 85±2 kJ mol?1.  相似文献   

19.
Absolute rate constants and their temperature dependence were determined by time-resolved electron spin resonance for the addition of the radicals Ph?H2 and Ph?Me2 to a variety of alkenes in toluene solution. To vinyl monomers CH2=CXY, Ph?H2 adds at the unsubstituted C-atom with rate constants ranging from 14 M ?1S ?1 (ethoxyethene) to 6.7 · 103 M ?1S ?1 (4-vinylpyridine) at 296 K, and the frequency factors are in the narrow range of log (A/M ?1S ?1) = 8.6 ± 0.3, whereas the activation energy varies with the substituents from ca. 51 kJ/mol to ca. 26 kJ/mol. The rate constants and the activation energies increase both with increasing exothermicity of the reaction and with increasing electron affinity of the alkenes and are mainly controlled by the reaction enthalpy, but are markedly influenced also by nucleophilic polar effects for electron-deficient substrates. For 1,2-disubstituted and trisubstituted alkenes, the rate constants are affected by additional steric substituent effects. To acrylate and styrenes, Ph?Me2 adds with rate constants similar to those of Ph?H2, and the reactivity is controlled by the same factors. A comparison with relative-rate data shows that reaction enthalpy and polar effects also dominate the copolymerization behavior of the styrene propagation radical.  相似文献   

20.
Modifying the β‐diketimine ligand LH 1 (LH=[ArN?C(Me)? CH?C(Me)? NHAr], Ar=2,6‐iPr2C6H3) through replacement of the proton in 3‐position by a benzyl group (Bz) leads to the new BzLH ligand 2, which could be isolated in 77 % yield. According to 1H NMR spectroscopy, 2 is a mixture of the bis(imino) form [(ArN?C(Me)]2CH(Bz) 2a and its tautomer [ArN?C(Me)? C(Bz)?C(Me)NHAr] 2b. Nevertheless, lithiation of the mixture of 2a and 2b affords solely the N‐lithiated β‐diketiminate [ArN?C(Me)? C(Bz)?C(Me)? NLiAr], BzLLi 3. The latter reacts readily with GeCl2?dioxane to form the chlorogermylene BzLGeCl 4, which serves as a precursor for a new zwitterionic germylene by dehydrochlorination with LiN(SiMe3)2. This reaction leads to the zwitterionic germylene BzL′Ge: 5 (BzL′=ArNC(?CH2)C(Bz)?C(Me)NAr) which could be isolated in 83 % yield. The benzyl group has a distinct influence on the reactivity of zwitterionic 5 in comparison to its benzyl‐free analogue, as shown by the reaction of 5 with phenylacetylene, which yields solely the 1,4‐addition product 6, that is, the alkynyl germylene BzLGeCCPh. Compounds 2–8 have been fully characterized by multinuclear NMR spectroscopy, mass spectrometry, elemental analyses, and single‐crystal X‐ray diffraction analyses.  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