共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
Rui‐Yan Li Bing‐Qiang Wang Zhi‐Ru Li Di Wu Ying Li 《International journal of quantum chemistry》2008,108(1):151-160
Using the counterpoise‐corrected potential energy surface method, the stationary structures of the π Br‐bond complexes C2H4‐nFn? BrF (n = 0–2) with all real frequencies have been obtained at MP2/aug‐cc‐pVDZ level. The order of the π Br‐bond length is 2.625 Å (C2H4? BrF) < 2.714 Å (C2H3F? BrF) < 2.751 Å (g‐C2H2F2? BrF) < 2.771 Å (trans‐C2H2F2? BrF) < 2.778 Å (cis‐C2H2F2? BrF). The interaction energies (Eint) are, respectively,‐5.9 (C2H4? BrF),‐4.4 (C2H3F? BrF),‐3.7 (g‐C2H2F2? BrF),‐3.1 (cis‐C2H2F2? BrF),‐2.8 kcal/mol (trans‐C2H2F2? BrF), at the CCSD (T)/aug‐cc‐pVDZ level, which include larger electron correlation contributions (Ecorre). The order of Ecorre is‐3.40 (C2H4? BrF),‐3.60 (C2H3F? BrF),‐3.85 (g‐C2H2F2? BrF),‐3.86 (cis‐C2H2F2? BrF),‐3.88 kcal/mol (trans‐C2H2F2? BrF). The earlier results show above that the F substituent effect elongates the π Br‐bond, reduces the Eint, and increases the Ecorre contribution of the interaction energy. Interestingly, the interaction energy of the cis‐C2H2F2? BrF structure with longer interaction distance is larger than that of the corresponding trans‐C2H2F2? BrF structure with shorter interaction distance. This reason comes from a special secondary interaction between lone pairs of Br atom with positive charge and some atoms (H, C) with positive charges of C2H2F2 in the cis‐C2H2F2? BrF structure. Comparing with corresponding C2H4‐nFn? ClF and C2H4‐nFn? HF, the C2H4‐nFn? BrF system has the larger Eint in which main contribution comes from the larger Ecorre, representing the larger dispersion interaction. The larger Ecorre contribution of the Eint of π Br‐bond can be used to understand that the π Br‐bond is shorter and stronger than corresponding π Cl‐bond. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2008 相似文献
3.
4.
5.
Pablo Morn‐Poladura Eduardo Rubio Jos M. Gonzlez 《Angewandte Chemie (Weinheim an der Bergstrasse, Germany)》2015,127(10):3095-3098
The cycloisomerization reaction of 1‐(iodoethynyl)‐2‐(1‐methoxyalkyl)arenes and related 2‐alkyl‐substituted derivatives gives the corresponding 3‐iodo‐1‐substituted‐1H‐indene under the catalytic influence of IPrAuNTf2 [IPr=1,3‐bis(2,6‐diisopropyl)phenylimidazol‐2‐ylidene; NTf2=bis(trifluoromethanesulfonyl)imidate]. The reaction takes place in 1,2‐dichloroethane at 80 °C, and the addition of ttbp (2,4,6‐tri‐tert‐butylpyrimidine) is beneficial to accomplish this new transformation in high yield. The overall reaction implies initial assembly of an intermediate gold vinylidene upon alkyne activation by gold(I) and a 1,2‐iodine‐shift. Deuterium labeling and crossover experiments, the magnitude of the recorded kinetic primary isotopic effect, and the results obtained from the reaction of selected stereochemical probes strongly provide support for concerted insertion of the benzylic C H bond into gold vinylidene as the step responsible for the formation of the new carbon–carbon bond. 相似文献
6.
《International journal of quantum chemistry》2018,118(19)
Sulfite reductase (SiR) catalyzes a six electron and six proton reduction of sulfite to sulfide. Similarly to the cytochrome P450 (cytP450) family, the active site in SiR contains a (partially reduced) heme bound axially to a cysteinate ligand—though with an extra Fe4S4 cluster. Fe(III) SO2−, Fe(III) SOH−, and Fe(III) SO(H2) intermediates have been proposed for the catalytic cycle of SiR, leading to a formally Fe(V)S species—akin to the widely accepted reaction mechanism in cytP450. Here, density functional theory (DFT) data is reported for of such FeSO(H2) intermediates. The Fe(III) SO2− models display relatively high energies for homolytic bond breaking compared to their isomeric oxygen‐bound Fe(III) OS2− models, and thus offer a better alternative in terms of avoiding radical side products able to induce enzyme suicide. This could be due to the fact that the (iron‐bound) sulfur is more active from a redox standpoint compared to oxygen, thus permitting the departing oxygen to maintain a redox‐inert state. Di‐protonation of the oxygen is computed to lead to a compound I type Fe(IV)S coupled to a porphyrin radical anion—consistent with an intermediate previously observed by x‐ray crystallography. 相似文献
7.
8.
9.
Ba2(Ni1?xLix)Ni2N2: A Low-Valency Nitridoniccolate with Puckered Layers [(NiN2/2)? (Ni1?xLix)? (NiN2/2)] Ba2(Ni1?xLix)Ni2N2 is obtained by reaction of lithium-barium-melts (molar ratios Li : Ba between 1 : 1 and 3 : 1) with nitrogen (1 atm.) in nickel-crucibles within a period of 15 h. Single crystals with a dark-metallic lustre are formed by cooling the melt to room temperature with a rate of 10°C/h (orthorhombic, Cmca; a = 713.3(2)pm, b = 1027.4(7)pm, c = 752.2(4)pm; z = 4; Dxr = 5.50 g/cm3 with x = 0.43). The crystal structure contains nearly liner [NiN2/2]-chains (N? Ni? N: 178.5(7)°, Ni? N? Ni 173.4(7)°; Ni? N: 178.6(1)pm), running parallel to the [100] direction, which are interconnected via (Ni1?xLix)-sites (linear units (N? (Ni1?xLix)? N); bond-lenths: 194.5(12)pm with x = 0.43) to form puckered layer [(Nin2/2)? (Ni1?xLix)? (NiN2/2)]. Barium is in a distorted trigonal-planar coordination by nitrogen atoms (Ba? N: 281.1(11)pm ? 285.5(11)pm. The nitrogen-coordination corresponds to a distorted octahedron, NBa3(Ni1?xLix)Ni2, with nickel in trans-position. The crystal structure of Ba2(Ni1?xLix)Ni2N2 is closely related to the Li3N-type structure: Li2[LiN] ? Ba{(Ni1?xLix)0.5 □ 0.5}[NiN]. Furthermore, this structure enlarges the scope of barium-nitrido-niccolates which up to now were found to contain merely [NiN2/2]-chains(BaNiN: Planar zigzag-chains; Ba8Ni6N7 helical zigzag-chains). 相似文献
10.
Chao Wang Changpeng Chen Jingyu Zhang Jian Han Qian Wang Kun Guo Pei Liu Mingyu Guan Yingming Yao Yingsheng Zhao 《Angewandte Chemie (Weinheim an der Bergstrasse, Germany)》2014,126(37):10042-10046
An easily synthesized and accessible N,O‐bidentate auxiliary has been developed for selective C H activation under palladium catalysis. The novel auxiliary showed its first powerful application in C H functionalization of remote positions. Both C(sp2) H and C(sp3) H bonds at δ‐ and ε‐positions were effectively activated, thus giving tetrahydroquinolines, benzomorpholines, pyrrolidines, and indolines in moderate to excellent yields by palladium‐catalyzed intramolecular C H amination. 相似文献
11.
David Hugas Dr. Sílvia Simon Dr. Miquel Duran Prof. Dr. Célia Fonseca Guerra Dr. F. Matthias Bickelhaupt Prof. Dr. 《Chemistry (Weinheim an der Bergstrasse, Germany)》2009,15(23):5814-5822
Dihydrogen bond or H 2 molecule? The central H? H bond in linear H4 can exist in two qualitatively different bonding modes corresponding to two different electronic states, namely a donor–acceptor dihydrogen bond (DHB) and a central H2 molecule with an electron‐pair bond (see figure). This insight evolves from Kohn–Sham density functional analysis and it is further applied here to understand the bonding in more realistic model systems.
12.
Sulfoximide and Sulfoximidium Salts – Structures and Hydrogen Bonding In the solid state dimethylsulfoximide ( 1 ) (orthorhombic; space group Pbca; a = 577.8, b = 931.2 and c = 1645.6 pm) makes intermolecular N? H ? N hydrogen bonds. The hydrogen halide salts (CH3)2S(O)NH2+Hal? (( 2 ), Hal??Cl?; ( 4 ), Hal??Br?) reacts with metal halides to yield (CH3)2S(O)NH2+MHal with the complex anions (( 5 ), MHal?SbCl4?; ( 6 ), MHal?SbCl52?; ( 7 ), MHal?SbCl6?; ( 8 ), MHal?SbBr52?; ( 9 ), MHal?AlCl4?). 2 crystallizes from ethanol (96%) as [(CH3)2S(O)NH2+Cl?]2 · H2O ( 3 ). The structures of 3 (monoclinic; space group P21/c; a = 917.0, b = 1344.7, c = 1080.8 pm and β = 103.8°; Z = 10), 4 (orthorhombic; space group Pbcn; a = 1028.9, b = 1132.6, c = 1074.1 pm; Z = 8) and 6 (monoclinic; space group C2/c; a = 2041.1, b = 1101.4, c = 3365.6 pm and β = 153.8°; Z = 8) are determined by X-ray analysis. In 6 Sb is coordinated in a distorted octahedra by 6 Cl in three short (mean 245,5 pm; SbCl3) and three long distances (291 to 299 pm; Cl?). Two of the chloride ions connect the Sb atoms to infinite Sb …? Cl …? Sb chains. Except for 7 and 9 there are bridges between the NH2 groups and the halide ions. The NH valence vibrations are discussed in view of hydrogen bonding. 相似文献
13.
Zhen Wang Jizhi Ni Yoichiro Kuninobu Motomu Kanai 《Angewandte Chemie (Weinheim an der Bergstrasse, Germany)》2014,126(13):3564-3567
The first copper‐catalyzed intramolecular C(sp3) H and C(sp2) H oxidative amidation has been developed. Using a Cu(OAc)2 catalyst and an Ag2CO3 oxidant in dichloroethane solvent, C(sp3) H amidation proceeded at a terminal methyl group, as well as at the internal benzylic position of an alkyl chain. This reaction has a broad substrate scope, and various β‐lactams were obtained in excellent yield, even on gram scale. Use of CuCl2 and Ag2CO3 under an O2 atmosphere in dimethyl sulfoxide, however, leads to 2‐indolinone selectively by C(sp2) H amidation. Kinetic isotope effect (KIE) studies indicated that C H bond activation is the rate‐determining step. The 5‐methoxyquinolyl directing group could be removed by oxidation. 相似文献
14.
15.
Jun Xing Zu Peng Chen Fang Yuan Xiao Xue Yan Ma Ci Zhang Wen Dr. Zhen Li Prof. Hua Gui Yang 《化学:亚洲杂志》2013,8(6):1265-1270
Multicomponent Cu? Cu2O? TiO2 nanojunction systems were successfully synthesized by a mild chemical process, and their structure and composition were thoroughly analyzed by X‐ray diffraction, transmission electron microscopy, field‐emission scanning electron microscopy, and X‐ray photoelectron spectroscopy. The as‐prepared Cu? Cu2O? TiO2 (3 and 9 h) nanojunctions demonstrated higher photocatalytic activities under UV/Vis light irradiation in the process of the degradation of organic compounds than those of the Cu? Cu2O, Cu? TiO2, and Cu2O? TiO2 starting materials. Moreover, time‐resolved photoluminescence spectra demonstrated that the quenching times of electrons and holes in Cu? Cu2O? TiO2 (3 h) is higher than that of Cu? Cu2O? TiO2 (9 h); this leads to a better photocatalytic performance of Cu? Cu2O? TiO2 (3 h). The improvement in photodegradation activity and electron–hole separation of Cu? Cu2O? TiO2 (3 h) can be ascribed to the rational coupling of components and dimensional control. Meanwhile, an unusual electron–hole transmission pathway for photocatalytic reactions over Cu? Cu2O? TiO2 nanojunctions was also identified. 相似文献
16.
Burkhard Butschke Maria Schlangen Detlef Schröder Helmut Schwarz 《Helvetica chimica acta》2008,91(10):1902-1915
Mechanistic details for the formation of methane from the title compound as well as the combined elimination of (CH3)2S/CH4 are derived from various mass‐spectrometric experiments including deuterium‐labeling studies and DFT calculations. For the first process, i.e., methane formation, we have identified three competing pathways in which the intact, Pt‐bonded methyl group combines with a H‐atom that originates from a phenyl substituent (ca. 7%), the dimethyl sulfide ligand (ca. 41%), and a methyl group of the diazabutadiene backbone (ca. 52%). In contrast, in the combined (CH3)2S/CH4 elimination, the methane is specifically formed from the Pt‐bound CH3 group and a H‐atom provided by one of the phenyl groups (‘cyclometalation’). 相似文献
17.
The conformational stability of aminomethanol and its methylated derivatives has been investigated by means of ab initio methods in the gas phase and aqueous solution. Among the computational levels employed, HF/6‐31G**//HF/6‐31G** calculations correctly describe the conformational features of this series of compounds, and agree well with the results obtained using larger basis sets and including ZPE or electron correlation corrections. Calculated energies and geometries follow the known trends associated to the generalized anomeric effect. Thus, the most stable conformers exhibit preferences for the trans orientations of the Lp N C O and Lp O C N moieties. However, reverse anomeric effects are observed when a methyl group is bonded to the oxygen, because the Lp O C N unit prefers a gauche orientation (that is, trans Me O C N). The natural bond orbital (NBO) method was employed to explain the cited conformational preferences. According to the NBO results, trans arrangements are preferred because the stabilization due to charge delocalization is more important than electrostatic and steric contributions. This explanation agrees with the conclusions obtained by other independent procedures based on energy decomposition schemes. The NBO method was also used to explain the origin of the rotational barriers around the C O and C N bonds in terms of the balance between unfavorable hyperconjugation and electrostatic and steric effects. Changes in conformational stability caused by methylations in different molecular positions were also explained by the influence of the methyl groups on lone‐pair delocalization and on steric effects. Finally, the effect of solvation was studied by means of the ab initio PCM method, and the significant changes on relative energies found were analyzed. © 2000 John Wiley & Sons, Inc. J Comput Chem 21: 462–477, 2000 相似文献
18.
Marko Rodewald Dr. J. Mikko Rautiainen Dr. Tobias Niksch Dr. Helmar Görls Dr. Raija Oilunkaniemi Prof. Dr. Wolfgang Weigand Prof. Dr. Risto S. Laitinen 《Chemistry (Weinheim an der Bergstrasse, Germany)》2020,26(61):13806-13818
The Te ⋅⋅⋅ Te secondary bonding interactions (SBIs) in solid cyclic telluroethers were explored by preparing and structurally characterizing a series of [Te(CH2)m]n (n=1–4; m=3–7) species. The SBIs in 1,7-Te2(CH2)10, 1,8-Te2(CH2)12, 1,5,9-Te3(CH2)9, 1,8,15-Te3(CH2)18, 1,7,13,19-Te4(CH2)20, 1,8,15,22-Te4(CH2)24 and 1,9,17,25-Te4(CH2)28 lead to tubular packing of the molecules, as has been observed previously for related thio- and selenoether rings. The nature of the intermolecular interactions was explored by solid-state PBE0-D3/pob-TZVP calculations involving periodic boundary conditions. The molecular packing in 1,7,13,19-Te4(CH2)20, 1,8,15,22-Te4(CH2)24 and 1,9,17,25-Te4(CH2)28 forms infinite shafts. The electron densities at bond critical points indicate a narrow range of Te ⋅⋅⋅ Te bond orders of 0.12–0.14. The formation of the shafts can be rationalized by frontier orbital overlap and charge transfer. 相似文献
19.
Dr. David E. Herbert Nadia C. Lara Prof. Theodor Agapie 《Chemistry (Weinheim an der Bergstrasse, Germany)》2013,19(48):16453-16460
The meta‐terphenyl diphosphine, m‐P2, 1 , was utilized to support Ni centers in the oxidation states 0, I, and II. A series of complexes bearing different substituents or ligands at Ni was prepared to investigate the dependence of metal–arene interactions on oxidation state and substitution at the metal center. Complex (m‐P2)Ni ( 2 ) shows strong Ni0–arene interactions involving the central arene ring of the terphenyl ligand both in solution and the solid state. These interactions are significantly less pronounced in Ni0 complexes bearing L‐type ligands ( 2‐L : L=CH3CN, CO, Ph2CN2), NiIX complexes ( 3‐X : X=Cl, BF4, N3, N3B(C6F5)3), and [(m‐P2)NiIICl2] ( 4 ). Complex 2 reacts with substrates, such as diphenyldiazoalkane, sulfur ylides (Ph2S?CH2), organoazides (RN3: R=para‐C6H4OMe, para‐C6H4CF3, 1‐adamantyl), and N2O with the locus of observed reactivity dependent on the nature of the substrate. These reactions led to isolation of an η1‐diphenyldiazoalkane adduct ( 2‐Ph2CN2 ), methylidene insertion into a Ni? P bond followed by rearrangement of a nickel‐bound phosphorus ylide ( 5 ) to a benzylphosphine ( 6) , Staudinger oxidation of the phosphine arms, and metal‐mediated nitrene insertion into an arene C? H bond of 1 , all derived from the same compound ( 2 ). Hydrogen‐atom abstraction from a NiI–amide ( 9 ) and the resulting nitrene transfer supports the viability of Ni–imide intermediates in the reaction of 1 with 1‐azido‐arenes. 相似文献
20.
Shi‐Liang Shi Stephen L. Buchwald 《Angewandte Chemie (Weinheim an der Bergstrasse, Germany)》2015,127(5):1666-1670
The synthesis of 3,3‐difluoro‐2‐oxindoles through a robust and efficient palladium‐catalyzed C H difluoroalkylation is described. This process generates a broad range of difluorooxindoles from readily prepared starting materials. The use of BrettPhos as the ligand was crucial for high efficiency. Preliminary mechanistic studies suggest that oxidative addition is the rate‐determining step for this process. 相似文献