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 31 毫秒
1.
李晓艳  孙政  孟令鹏  郑世钧 《化学学报》2007,65(20):2203-2210
利用量子化学从头算CASSCF方法在6-311+G (d, p)基组水平上对单线态和三线态RN (R=CH3, CH3CH2)异构化反应及RN脱氢反应的微观机理进行了理论研究. 在MP2/6-311+G (d, p)和CCSD/6-311+G (d, p)水平上进行了单点能校正. 单态和三态势能面的交叉点(ISC)的存在清楚地说明了基态反应物3RN异构化为基态产物1R'NH (R'=CH2, CH3CH)的过程. 电子密度拓扑分析显示在整个异构化过程中有两种类型的结构过渡态: 单态反应通道为T型过渡态, 三态反应通道为环状过渡态. 单线态RN脱氢反应通道中“原子-分子键”的存在说明两个H原子是以H2的形式从RN中脱去的.  相似文献   

2.
A detailed investigation has been performed at the QCISD(T)/6‐311++G(d,p)//B3LYP/6‐311+G(d,p) level for the reaction of NCO with C2H5 by constructing singlet and triplet potential energy surfaces (PES). The results show that the title reaction is more favorable on the singlet PES than on the triplet PES. On the singlet PES, the initial addition processes are barrierless and release lots of energy. The dominant channel occurs via the fragmentations of the initial adduct C2H5NCO and C2H5OCN to form C2H4 + HNCO and HOCN, respectively. With higher barrier heights, other products such as CH4 + HNC + CO, CH3CHNH + CO, CH3CH + HNCO, and CH3CN + H2 + CO are less competitive. On the triplet PES, the entrance reactions surpass significant barriers; therefore, it could be negligible at the normal atmospheric condition. However, the most feasible channel on the triplet PES is the direct hydrogen abstraction channel to form CH2CH2 + HNCO. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

3.
Two new uranyl complexes [UO2(DPDPU)2(NO3)2](C6H5CH3) (1) and [UO2(PMBP)2 (DPDPU)](CH3C6H4CH3)0.5 (2), (DPDPU?=?N,N′-dipropyl-N,N′-diphenylurea, HPMBP?= 1-phenyl-3-methyl-4-benzoyl-pyrazolone-5) were synthesized and characterized. The coordination geometry of the uranyl atom in 1 is distorted hexagonal bipyramidal, coordinated by two oxygen atoms of two DPDPU molecules and four oxygen atoms of two bidentate nitrate groups. The coordination geometry of the uranyl atom in 2 is distorted pentagonal bipyramidal, coordinated by one oxygen atom of one DPDPU molecule and four oxygen atoms of two chelating PMBP molecules.  相似文献   

4.
Two new tetra‐ or di‐α‐substituted zinc(II) phthalocyanines 5 and 6 have been prepared through a “side‐strapped” method. In the molecules, the adjacent benzene rings of the phthalocyanine core are linked at α‐position through a triethylene glycol bridge to form a hybrid aza‐/oxa‐crown ether. The tetra‐α‐substituted phthalocyanine 5 shows an eclipsed self‐assembly property in CH2Cl2 and the effect on the di‐α‐substituted analogue 6 is significantly weakened. Furthermore, the crown ethers of these compounds can selectively complex with Fe3+ or Cu2+ ion in DMF, leading to formation of J‐aggregated nano‐assemblies, which can be disaggregated in the presence of some organic or inorganic ligands, such as triethylamine, tetramethylethylenediamine, CH3COO?, or OH?. In addition, both compounds are efficient singlet oxygen generators with the singlet oxygen quantum yields (ΦΔ) of 0.54‐0.74 in DMF relative to unsubstituted zinc(II) phthalocyanine (ΦΔ=0.56). They exhibit photodynamic activities toward HepG2 human hepatocarcinoma cells, but the compound 6 , which has more than 40‐fold lower IC50 value (0.08 μM ) compared to the analogue 5 (IC50=3.31 μM ), shows remarkablely higher in vitro photocytotoxicity due to its significantly higher cellular uptake and singlet oxygen generation efficiency. The results suggest that these compounds can serve as promising multifunctional materials both in (opto)electronic field and photodynamic therapy.  相似文献   

5.
Rose bengal-sensitized photooxygenation of 4-propyl-4-octene ( 1 ) in MeOH/Me2CHOH 1:1 (v/v) and MeOH/H2O 95:5 followed by reduction gave (E)-4-propyl-5-octen-4-ol ( 4 ), its (Z)-isomer 5 , (E)-5-propyl-5-octen-4-ol ( 6 ), and its (Z)-isomer 7 . Analogously, (E)-4-propyl[1,1,1-2H3]oct-4-ene ( 2 ) gave (E)-4-propyl[1,1,1-2H3]oct-5-en-4-ol ( 14 ), its (Z)-isomer 15 , (E)-5-[3′,3′,3′-2H3]propyl-5-octen-4-ol ( 16 ), its (Z)-isomer 17 , and the corresponding [8,8,8-2H3]-isomers 18 and 19 (see Scheme 1). The proportions of 4–7 were carefully determined by GC between 10% and 85% conversion of 1 and were constant within this range. The labeled substrate 2 was photooxygenated in two high-conversion experiments, and after reduction, the ratios 16/18 and 17/19 were determined by NMR. Isotope effects in 2 were neglected and the proportions of corresponding products from 1 and 2 assumed to be similar (% 4 ≈? % 14 ; % 5 ≈? % 15 ; % 6 ≈? % ( 16 + 18 ): % 7 ≈? % ( 17 + 19 )). Combination of these proportions with the ratios 16/18 and 17/19 led to an estimate of the proportions of hydroperoxides formed from 2 . Accordingly, singlet oxygen ene additions at the disubstituted side of 2 are preferred (ca. 90%). The previously studied trisubstituted olefins 20–25 exhibited the same preference, but had both CH3 and higher alkyl substituents on the double bond. In these substrates, CH3 groups syn to the lone alkyl or CH3 group appear to be more reactive than CH2 groups at that site beyond a statistical bias.  相似文献   

6.
The results obtained from CASSCF‐MRMP2 calculations are used to rationalize the singlet complexes detected under matrix‐isolation conditions for the reactions of laser‐ablated Zr(3F) atoms with the CH3F and CH3CN molecules, without invoking intersystem crossings between electronic states with different multiplicities. The reaction Zr(3F) + CH3F evolves to the radical products ZrF· + ·CH3. This radical asymptote is degenerate to that emerging from the singlet channel of the reactants Zr(1D) + CH3F because they both exhibit the same electronic configuration in the metal fragment. Hence, the caged radicals obtained under cryogenic‐matrix conditions can recombine through triplet and singlet paths. The recombination of the radical species along the low‐multiplicity channel produces the inserted structures H3C? Zr? F and H2C?ZrHF experimentally detected. For the Zr(3F) + CH3CN reaction, a similar two‐step reaction scheme involving the radical fragments ZrNC· + ·CH3 explains the presence of the singlet complexes H3C? Zr? NC and H2C?Zr(H)NC revealed in the IR‐matrix spectra upon UV irradiation. © 2014 Wiley Periodicals, Inc.  相似文献   

7.
Eight coumarins, which carry a terminal alkene tethered by a CH2XCH2 group to their 4‐position (X=CH2, CMe2, O, S, NBoc, NZ, NTs, NBn), were synthesized in overall yields of 51–80 %. Starting materials for the syntheses were either commercially available 4‐hydroxycoumarin or 4‐formylcoumarin. The intramolecular [2+2] photocycloaddition of these coumarins gave diastereoselectively products with a tetracyclic 3,3a,4,4a‐tetrahydro‐1H‐cyclopenta[2,3]cyclobuta[1,2‐c]chromen‐5(2H)‐one skeleton. Direct irradiation at λ=300 nm in dichloromethane (c=10 mM ) led to product formation in good yields for most substrates, presumably via a singlet excited state intermediate. Due to the low coumarin absorption at λ >350 nm the photocycloaddition was slow upon irradiation at λ=366 nm. Addition of a chiral oxazaborolidine‐based Lewis acid (50 mol %) increased the reaction rate at λ=366 nm and induced a significant enantioselectivity in the [2+2] photocycloaddition. Six out of eight coumarin substrates (X=CH2, CMe2, O, NBoc, NZ, NTs) gave the respective products in yields of 72–96 % and with 74–90 % enantiomeric excess (ee) upon irradiation in dichloromethane (c=20 mM ) at ?75 °C. The Lewis acid presumably acts by coordination to the coumarin carbonyl oxygen atom, which leads to a bathochromic shift (redshift) of the UV absorption and which increases the singlet state lifetime. A second electrostatic interaction of the hydrogen atom at C3 with the oxygen atom of the oxazaborolidine is likely.  相似文献   

8.
Synthetic routes for the preparation of 3-alkyl-6-phenyl-4(3H)-pteridinones 6 and their corresponding 8-oxides 5 (R = CH3, C2H5, (CH2)2CH3, (CH2)3CH3, CH(CH3)C2H5, CH(CH3)2 and CH(C2H5)CH2OCH(OC2H5)2 are described and their reactivities towards xanthine oxidase from Arthrobacter M-4 are determined. Only the 3-methyl derivative of 6-phenyl-4(3H)-pteridinone and its 8-oxide i. e. 6a and 5a are found to be substrates although their reactivities are still very low. Oxidation takes place at C-2 of the pteridinone nucleus. All the 3-alkyl derivatives are less tightly bound to the enzyme than 6-phenyl-4(3H)-pteridinone. Introduction of the N-oxide at N-8 considerably lowers the binding of the substrates. Inhibition studies have revealed that 3-methyl-6-phenyl-4(3H)-pteridinone ( 6a ) is a non-competitive inhibitor with a Ki-value of 47 μM and the 3-ethyl derivative ( 6b ) an uncompetitive one with a Ki-value of 19.6 μM.  相似文献   

9.
High-temperature (>1000°K) pyrolysis of acetaldehyde (~1% in an atmosphere of pure nitrogen) was examined in a turbulent flow reactor which permits accurate determination of the spatial distribution of the stable species. Results show that the products in order of decreasing importance are CO, CH4, H2, C2H6, and C2H4. Rates of formation were consistent with the Rice–Herzfeld mechanism by including reactions to explain C2H4 formation and the possible presence of ketene. A steady-state treatment of the complete mechanism indicates that the overall reaction order decreases from \documentclass{article}\pagestyle{empty}\begin{document}$ \frac{3}{2} $\end{document} to 1, which is supported by the new experimental data. Using earlier low-temperature results, the rate constant for the reaction CH3CHO → CH3 + CHO (1) was found as k1=1015.85±0.21 exp (?81,775±1000/RT) sec?1. Also, data for the ratio of rate constants for reactions CH3CHO + CH3 → CH4 + CH3CO (4) and 2CH3 → C2H6(6) were fitted to the empirical expression k4/k61/2=10?13.89±0.03T6.1 exp(?1720±70/RT) (cm3/mole·sec)1/2 and causes for the curvature are discussed. The noncatalytic effect of oxygen on acetaldehyde pyrolysis at high temperature is explained.  相似文献   

10.
尹汉东  薛绳才  王其宝 《中国化学》2004,22(10):1187-1191
Introduction Dimeric tetraorganodistannoxanes are a kind of in-teresting organotin oxo clusters and have attracted con-siderable attention during the last several decades, in view of their unique structural features1-5 as well as their applications as biocides6,7 and in homogenous cataly-sis.8,9 In the solid state, they contain characteristic Sn4O2X2Y2 structural motifts with staircase or ladder arrangements, a planar four-membered Sn2O2 ring and, generally, penta-coordination around the tin…  相似文献   

11.
The photo-oxygenation of adamantylideneadamantane ( 1 ) on siliceous supports using admixed granules of ion-exchange resin fixed to methylene blue (MB) and rose bengal (RB) gave exclusively the corresponding dioxetane derivative 2 for the former sensitizer, while the latter gave 2 and traces of the epoxide 3. RB and the charge-transfer complex produced from N-ethylcarbazole and 2,4,5,6-tetranitrofluoren-9-one both reacted with chemically generated singlet oxygen to give superoxide radical anion. Trapping of the latter with 5,5-dimethyl-1-pyrroline 1-oxide gave an adduct exhibiting a characteristic ESR spectrum. The treatment of 1 in MeOH with 30% aqueous H2O2 for 22 h at 60° gave 3 in 100% yield. Repetition of this experiment in the presence of 2,6-di(tert-butyl)-p-cresol caused no significant change. These results indicate that singlet oxygen reacts with 1 , in the presence of RB, by two different processes. The first leads to dioxetane formation. The second process involves conversion of singlet oxygen by RB to superoxide radical anion which subsequently gives H2O2 so producing epoxide 3 from 1 .  相似文献   

12.
Three flavonoid copper(II) complexes Cu2(quercetin)(CH3COO)3(CH3OH) ( 1 ), Cu(anthrarufin)(CH3COO)·1/2H2O ( 2 ) and Cu(naringin)(OCH3)(CH3OH)2 ( 3 ) have been synthesized and characterized by elemental analysis, IR, electronic absorption and EPR (X‐band) spectroscopy. The complexes have a strong protective action over the Δsod1 mutant of S. cerevisiae against reactive oxygen radicals generated by an external source of free radicals (H2O2 or the superoxide‐generating, menadione). On the other hand, the complexes cleave DNA efficiently even in the absence of reducing agents. The main reactive oxygen species responsible for the DNA strand cleavage have been determined using radical scavengers. A probably mechanism of the DNA damage is proposed.  相似文献   

13.
Two new divalent metal aminodiphosphonates with a layered structure, Mn2[(HL)(H2O)F]·H2O (1) and Cd3.5[(HL)L] (2) (H4L =?CH3CH2CH2N(CH2PO3H2)2), have been hydrothermally synthesized and characterized by single-crystal X-ray diffraction as well as with infrared spectroscopy, elemental analysis and thermogravimetric analysis. In 1, two MnO4F2 and two MnO5F polyhedra are interconnected via edge-sharing into a tetramer, and such tetramers are bridged by the diphosphonate ligands into a Mn(II) phosphonate layer in the ab-plane. The structure of 2 also features a 2D layered structure, in which the CdO5N and CdO6 polyhedra are interconnected into a 1D chain. The chains are then cross-linked via phosphonate oxygen atoms to form Cd(II) phosphonate layers in the ab-plane.  相似文献   

14.
Effects of steric crowding of the substituent of carboxylate counteranions on living cationic polymerization of isobutyl vinyl ether (IBVE) were investigated with the use of two series of carboxylic acids with various carbonyl substituents [RCOOH; R = (aliphatic series) CH3CH2, (CH3)2CH, (CH3)3C; (aromatic series) C6H5CH2, (C6H5)2CH, (C6H5)3C] in conjunction with tin tetrabromide (SnBr4) and 1,4-dioxane (DO) in toluene at 0°C. The overall polymerization rate increased with increasing the bulkiness of the substituents R in both the series: R = CH3 (1) ≃ CH3CH2 (1) < (CH3)2CH (1.76) < (CH3)3C (2.31); C6H5CH2 (0.84) < (C6H5)2CH (0.98) < (C6H5)3C (1.74); the values in the parentheses show the relative polymerization rate. In all the polymerizations, the number-average molecular weight (Mn) of the polymers was directly proportional to monomer conversion and in good agreement with the calculated values, assuming that one RCOOH molecule forms one polymer chain. The living nature of these polymerizations was further confirmed by a linear increase in Mn of the polymers upon sequential addition of a fresh monomer feed to the almost completely polymerized reaction mixtures. In the polymerizations with sterically less hindered carboxylic acids [R = CH3CH2, (CH3)2CH, C6H5CH2, (C6H5)2CH], the molecular weight distribution (MWD) of the polymers was very narrow (Mw/Mn < 1.1) throughout the polymerizations. In contrast, with bulkier substituent-containing counterparts [R = (CH3)3C, (C6H5)3C], the polymerizations led to the polymers of relatively broad MWD (Mw/Mn ≅ 1.5 at ca. 100% monomer conversion). The bulky substituents such as (CH3)3C and (C6H5)3C may decrease the interconversion rate between a dormant and an active species and increase the time-average concentration of the active growing species. The stereoregularity of the obtained polymers was not changed much with the steric environment of the counteranion (meso: 66–69%). © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 2923–2932, 1999  相似文献   

15.
Ab initio and density functional CCSD(T)-F12/cc-pVQZ-f12//B2PLYPD3/6-311G** calculations have been performed to unravel the reaction mechanism of triplet and singlet methylene CH2 with ketene CH2CO. The computed potential energy diagrams and molecular properties have been then utilized in Rice–Ramsperger–Kassel–Marcus-Master Equation (RRKM-ME) calculations of the reaction rate constants and product branching ratios combined with the use of nonadiabatic transition state theory for spin-forbidden triplet-singlet isomerization. The results indicate that the most important channels of the reaction of ketene with triplet methylene lead to the formation of the HCCO + CH3 and C2H4 + CO products, where the former channel is preferable at higher temperatures from 1000 K and above. In the C2H4 + CO product pair, the ethylene molecule can be formed either adiabatically in the triplet electronic state or via triplet-singlet intersystem crossing in the singlet electronic state occurring in the vicinity of the CH2COCH2 intermediate or along the pathway of CO elimination from the initial CH2CH2CO complex. The predominant products of the reaction of ketene with singlet methylene have been shown to be C2H4 + CO. The formation of these products mostly proceeds via a well-skipping mechanism but at high pressures may to some extent involve collisional stabilization of the CH3CHCO and cyclic CH2COCH2 intermediates followed by their thermal unimolecular decomposition. The calculated rate constants at different pressures from 0.01 to 100 atm have been fitted by the modified Arrhenius expressions in the temperature range of 300–3000 K, which are proposed for kinetic modeling of ketene reactions in combustion. © 2018 Wiley Periodicals, Inc.  相似文献   

16.
The reaction of nickelocene with BrMgR, where R=CH2CH(CH3)C6H5, C2H5, (CH2)7CH3 and CH2CH2CH3, have been studied. It was found that the presence of β-hydrogen in R did not cause the total splitting of the carbon–nickel bond but alkylidynetrinickel clusters were formed. It is the first example of the synthesis of alkylidynetrinickel clusters (NiCp)3CR′ from the organonickel species possessing β-hydrogen. Besides trinickel clusters, the following compounds were always formed in all the studied reactions: (NiCp)4H2, (NiCp)6, CpNi(η3-C5H7) and (NiCp)2(μ-C5H6). The structure of (NiCp)3CCH(CH3)Ph has been determined by a single-crystal X-ray diffraction study.  相似文献   

17.
The radical-molecule reaction mechanisms of CH2Br and CHBrCl with NO2 have been explored theoretically at the UB3LYP/6-311G(d, p) level. The single-point energies were calculated using UCCSD(T) and UQCISD(T) methods. The results show that the title reactions are more favorable on the singlet potential energy surface than on the triplet one. For the singlet potential energy surface of CH2Br + NO2 reaction, the association of CH2Br with NO2 is found to be a barrierless carbon-to-oxygen attack forming the adduct IM1 (H2BrCONO-trans), which can isomerize to IM2 (H2BrCNO2), and IM3 (H2BrCONO-cis), respectively. The most feasible pathway is the 1, 3-Br shift with C–Br and O–N bonds cleavage along with the N–Br bond formation of IM1 lead to the product P1 (CH2O + BrNO) which can further dissociate to give P4 (CH2O + Br + NO). The competitive pathway is the 1, 3-H-shift associated with O–N bond rupture of IM1 to form P2 (CHBrO + HNO). For the singlet potential energy surface of CHBrCl + NO2 reaction, there are three important reaction pathways, all of which may have comparable contribution to the reaction of CHBrCl with NO2. The theoretically obtained major products CH2O and CHClO for CH2Br + NO2 and CHBrCl + NO2 reactions, respectively, are in good agreement with the kinetic detection in experiment.  相似文献   

18.
The reaction of cis-(CO)4Fe[Si(CH3)3]2 (I) with CH3OSi(CH3)3 and C6H5CH2-OSi(CH3)3 at 80°C affords good yields of [(CH3)3Si]2O and the deoxygenation products RSi(CH3)3 (R = CH3, C6H5CH2). These reactions are proposed to occur via (CO)4Fe(R)Si(CH3)3 intermediates. This is supported by the observed formation of cis-(CO)4Fe(CH3)Si(CH3)3 (II) during the more rapid reaction of I with (CH3)2O; subsequent (CH3)4Si elimination occurs. With (C6H5CH2)2O, I reacts at 80°C to yield C6H5CH2Si(CH3)3 and C6H5CH2OSi(CH3)3 as primary products. With C6H5CH2OCH3, I effects regioselective benzyl---oxygen bond cleavage.  相似文献   

19.
Twelve new germanium substituted diphenyltin dipropionates with the general formula (R1GeCHR2‐CHR3COO)2SnPh2 where R1 = N(CH2CH2O)3, (C6H5)3 and (CH3C6H4)3, R2 = H, CH3, C6H5, p‐CH3C6H4, p‐CH3OC6H4, p‐ClC6H4, and R3 = H, CH3 have been synthesized by the reaction of diphenyltin oxide with a germanium substituted propionic acid. All the compounds were characterized by elemental analysis, IR, multi‐nuclear (1H, 13C, 119Sn) NMR and Mössbauer spectroscopies as well as mass spectrometry. The in vitro antibacterial activity of selected compounds is also reported.  相似文献   

20.
Specific magnetic susceptibilities (s) of several newly synthesized chelates of some of the lanthanons [La(III), Pr(III) and Nd(III)] are reported. These derivatives are of the general type,Ln(O-i-C3H7)3–n (C6H5CHNRO) n [where,Ln=La(III), Pr(III) or Nd(III);n=1 or 2 and R=CH2CH2, CH2CHCH3 or C6H4] and have been prepared by the reaction of the alkoxides of the lanthanons withSchiff bases such as benzylidene-2-hydroxyethylamine (C6H5CHNCH2CH2OH), benzylidene-2-hydroxy-n-propylamine (C6H5CHNCH2CHOHCH3) and benzylidene-o-aminophenol (C6H5CHNC6H4OH) in different molar relations in dry benzene.The resulting crystalline derivatives are non-volatile, light to deep yellow or blackish in colour. These tend to polymerize on keeping as shown by their insoluble nature and higher melting points, the polymerisation possibly occurring by the intermolecular coordination through oxygen atoms as reported earlier1.UsingGouy method2, the bis-isopropoxy mono-Schiff base and mono-isopropoxy bis-Schiff base complexes of La(III) have been shown to be diamagnetic, with s values being in the range of –0.32 to –0.45×10–6 and –0.39 to –0.55×10–6 c.g.s. units at 305 K respectively.In the remaining derivatives, Pr(O-i-C3H7)3–n (C6H5CH NRO) n and Nd(O-i-C3H7)3–n (C6H5CHNRO) n (where,n=1 or 2 and R=CH2CH2, CH2CHCH3 or C6H4) the magnetic moment values range between 3.25 to 3.32 and 3.30 to 3.33 B respectively indicating their paramagnetic nature.  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