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 62 毫秒
1.
The He(I) photoelectron spectrum of the N3(X 2H) radical, produced by the rapid gas phase reaction of fluorine atoms with hydrazoic acid, has been recorded. Three bands were observed corresponding to ionization of N3(X 2H) to the N3 + X 3Σ?, 1Δ and 1Σ+ states. The vertical ionization energies to the observed ionic states were compared with those obtained from ΔSCF ab initio calculations. Evidence is presented to show that the major pathway of the F+HN3 reaction occurs via a long-lived intermediate and involves abstraction of a hydrogen atom yielding vibrationally excited HF and the N3 radical. Consideration of both the photoelectron data and a recent chemiluminescence study of the reaction has led to some re-interpretation of the latter results.  相似文献   

2.
We have investigated the photoinduced decomposition of formaldehyde (CH2O) on a rutile TiO2(100)-(1×1) surface at 355 nm using temperature-programmed desorption. Products, formate (HCOO-), methyl radical (CH3·), ethylene (C2H4), and methanol (CH3OH) have been detected. The initial step in the decomposition of CH2O on the rutile TiO2(100)-(1×1) surface is the formation of a dioxymethylene intermediate in which the carbonyl O atom of CH2O is bound to a Ti atom at the five-fold-coordinated Ti4+ (Ti5c) site and its carbonyl C atom bound to a nearby bridge-bonded oxygen (Ob) atom, respectively. During 355 nm irradiation, the dioxymethylene intermediate can transfer an H atom to the Ob atom, thus forming HCOO- directly, which is considered as the main reaction channel. In addition, the dioxymethylene intermediate can also transfer methylene to the Ob row and break the C-O bond, thus leaving the original carbonyl O atom at the Ti5c site. After the transfer of methylene, several pathways to products are available. Thus, we have found that Ob atoms are intimately involved in the photoinduced decomposition of CH2O on the rutile TiO2(100)-(1×1) surface.  相似文献   

3.
Density functional theory calculations were performed for the methanolysis reactions of a set of aryloxy and alkoxy acetates ( 1a , 1b , 1c , 1d , 1e , 1f , 1g , 1h , 1i , 1j , 1k , 1l , 1m ) promoted by methoxide and a 1,5,9‐triazacyclododecane‐complexed Zn(II)‐methoxide [2(OCH3)]+ in order to give free energies and structural data for the various intermediates and transition states along the reaction pathway. The methoxide‐promoted reactions experience a transition of pathways from enforced‐concerted addition of CH3O? to the C = O unit for substrates having a good aryloxy leaving groups (LGs) with strong electron withdrawers ( 1a , 1b , 1c , 1d , 1e ) to a two step process with rate‐limiting CH3O attack on aryloxy acetates having higher (the pKa of the parent phenol of the LG in methanol) values. Only in the case of the substrates 1i‐m having alkoxy LGs is there an observed change in rate‐limiting step that occurs at the quasi‐symmetrical point where the . The methanolysis process for the 2,4‐dinitrophenoxy substrate ( 1a ) promoted by [2(OCH3)]+ involves transient binding of the substrate to the metal complex followed by a rate‐limiting, enforced‐concerted attack of Zn(II)‐coordinated OCH3, with fast breakdown of an addition intermediate that does not have a significant lifetime. For substrates 1b,c having slightly less electron withdrawing substituents, the reaction has two steps with rate‐limiting attack and an unassisted LG departure. As the increases, the reaction still has two steps with rate‐limiting attack, but departure of the LG is now assisted by its coordination to the metal ion. For alkoxy containing substrates, a change in rate‐limiting step occurs centered at methoxy acetate, 1j , (when ) for which the second step of metal ion assisted departure of methoxide becomes partially rate‐limiting. The Brønsted plots computed for the methoxide‐promoted and [2(OCH3)]+‐promoted methanolyses are compared with the previously determined experimental data and are analyzed as arising not from a common line attributable to all substrates but rather in terms of separate, but intersecting, plots for aryl‐ and alkyl acetates. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

4.
Contrary to the cleavage of α‐phenylthioureido peptides 1 proceeding through intermediate 2‐anilinothiazolinone 2 , the b‐analog cis‐2‐(3‐phenylthioureido)cyclopentane‐carboxamide 5 forms transiently 4‐imino‐2‐thioxopyrimidine 6 . Monitoring amide cyclization and hydrolysis of iminopyrimidine 6 in acid by UV showed that an equilibrium between 5 and 6 was reached followed by slower conversion of both compounds into 2‐oxo‐4‐thioxopyrimidine 7 . Both processes were characterized by isosbestic points, the first due to parallel conversion of 5 into 6 and 7 (or 6 into 5 and 7 ) at a constant ratio while the second identical for both reactants – to conversion of equilibrated 5 and 6 into 7 . The special isosbestic points allowed the determination of the individual constants of Scheme 2. Further confirmation was obtained from NMR product analysis and following the cyclization of amide 5 in DMSO:DCl. Product 2‐oxo‐4‐ thioxopyrimidine 7 hydrolyzed reversibly to thioureido acid 8 . The cyclization rate of 8 allowed the participation of 6‐oxothiazine 10 formed by sulfur attack to be excluded. The absence of sulfur attack in the six‐membered case is explained by the longer C? S bond bringing about greater bond angle strain at the tetrahedral ring atoms due to the geometrical characteristics of five‐ and six‐membered rings with planar segments. The cyclizations of amide 5 to iminopyrimidine 6 and to thiodihydrouracil 7 are first order in [H+], while the reactions of protonated imine 6 H+ are zero order to amide and ?1 to thiodihydrouracil. The reaction orders can be reconciled by assuming a rate determining proton transfer from the tetrahedral intermediate in amide cyclization. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

5.
Through complete neglect of differential overlap (CNDO) calculations of the electronic energy among different possible structures of paracetamol (PA) molecule, it has been concluded that its structure has Cs point group symmetry of the cis‐form in which the methyl group has a restricted free rotation around its bond with the carbon atom of the amide group. The electronic spectra of PA compound were studied in different polar and nonpolar solvents. The temperature effect on the electronic spectra confirms the presence of one conformer only. The hydrogen bonding and the orientation energies of the polar solvents were determined from the studies of mixed solvents. Complexes of PA with metal ions M(II) (Cu++, Zn++, or Fe++) of ratio 2:1, respectively, were prepared, and their structure has been confirmed by elemental analysis, atomic absorption spectra, IR spectra, and 1H‐NMR spectra. It has been concluded that the structure of the complexes has C2h point group symmetry in which two PA molecules are chelated to any one of the metal ions Cu++, Zn++, and Fe++.  相似文献   

6.
In this article, Zn3(PO4)2: Eu3+ nanorods and nanoparticles have been prepared by the hydrothermal method. The optimum pH value has been found at the range of 3–8 for the preparation of orthorhombic Zn3(PO4)2: Eu3+, whose morphologies are affected by the pH value. At the same temperature for hydrothermal reaction, the product presents nanorods at pH 4, while it shows nanoparticles at pH 6. Furthermore, the influences of the hydrothermal reaction temperature on the morphology and microstructure have also been investigated, suggesting that the morphology and microstructure cannot be changed with the hydrothermal temperature at the same pH value. Finally, the photoluminescence of Eu3+ on Zn3(PO4) nanorod/nanoparticle have been studied, both of which present the characteristic emission lines of Eu3+ and the 5D07F1 transition corresponds the strongest emission. This indicates that Eu3+ occupied the inversion center in Zn3(PO4) host.  相似文献   

7.
用密度泛函理论在B3LYP/6-31G(d,p)计算水平下研究了次氯酸钠氧化邻硝基苯胺生成苯并氧化呋咱的环氧化反应.考虑溶剂化效应对反应的影响,使用极化连续反应场模型进行几何优化.计算了该反应的两种可能反应通道,它们都是分步反应,反应通道A经历氧化、移氢、脱水和环化四步反应,在反应通道B中,氢氧化钠的OH-首先进攻邻硝基苯胺的胺基H原子,生成邻硝基苯亚胺负离子.计算结果表明,在反应通道A是可行的反应通道,1个水分子辅助进行分子内脱水反应是速控步骤.  相似文献   

8.
Ethyl formate undergoes spontaneous hydrolysis through two parallel mechanistic pathways: water catalysis and specific hydrogen-ion catalysis. The proton inventory of the water reaction at 37.0 °C was determined from the overall conversion curves, in H2O/D2O mixtures of deuterium atom fraction n, resulting in k 1(n)=(2.1±0.1)×10?5 s?1 (1?n+0.42 n) (1?n+0.83 n)2, in agreement with the results of six other cases of carboxylic acid derivatives. This result contributes to the validation of the accuracy of the proton inventory technique.  相似文献   

9.
The reaction of nitrogen dioxide with carbonyl compounds having a methylene or vinyl group adjacent to the corbonyl leads to the formation of stable iminoxy radicals1-3. Evidence1,2,4 has been presented that the precursor of these iminoxy radicals is probably the corresponding oxime formed by rearrangements of the intermediate nitroso compounds. There have also been reports on the e.s.r. detection of some cis-trans isomers of some iminoxy radicals in methanol or in neat liquid1,5. We wish to report here a drastic solvent effect on the conformations of some iminoxy radicals formed by the reaction of NO2 with trifluoroacetylacetone.  相似文献   

10.
Zn2?2x Mn2x GeO4 (x=0, 0.001, 0.01) phosphors were prepared by conventional solid state reaction technique. X-ray powder diffraction (XRD), field emission scanning electron microscopy (FE-SEM), diffuse reflection spectra, photoluminescence (PL), and cathodoluminescence (CL) spectroscopy were utilized to characterize the synthesized phosphors. The Mn2+-activated Zn2GeO4 phosphors exhibit narrow emission band at 532 nm under the excitation of ultraviolet light, which due to the 4T1(4G)–6A1(6S) transition of Mn2+ ions. Also it is observed that there exists energy transfer between the Zn2GeO4 host lattice and the activator (Mn2+). Under excitation of low-voltage electron beams, Zn2GeO4:Mn2+ shows strong green emission band dominating at 535 nm, corresponding to the 4T1(4G)–6A1(6S) emission of Mn2+ ions. The emission intensity and chromaticity coordinates of Zn2GeO4:Mn2+ as a function of accelerating voltage and the filament current were also investigated.  相似文献   

11.
The gas-phase thermal dehydration mechanism of tert-butanol, 2-methyl-2-butanol, 2-methyl-2-pentanol and 2,3-dimethyl-2-butanol by homogeneous catalysis of hydrogen bromide was examined by density functional theory calculations with the hybrid functionals: M062X, CAMB3LYP and WB97XD. Reasonable agreements were found between theoretical and experimental enthalpy values at the WB97XD/6-311++G(d,p) level. The dehydration mechanism of tert-butanol with and without catalysis was evaluated in order to examine the catalyst effect on the mechanism. The elimination reaction without catalysis involves a four-membered transition state (TS), while the reaction with catalysis involves a six-membered TS. The mechanism without catalysis has enthalpy activation over 150 kJ mol–1 greater than the catalysed reaction. In all these reactions, the elongation of the C–O bond is significant in the TS. The un-catalysed reaction is controlled by breaking of C–O bond, and it was found to be more synchronous (Sy ≈ 0.91) than the hydrogen bromide catalysed reactions (Sy ≈ 0.75–0.78); the latter reactions are dominated by the three reaction coordinates associated with water formation. No significant effect on the enthalpies of activation was observed when the size of the alkyl chain was increased.  相似文献   

12.
We have performed density functional theory (DFT) calculations in order to study the gas‐phase interaction of oxo‐ and thio‐oxazolidine derivatives with Zn2+. The calculations were performed at B3LYP/6‐311+(2df,2p) level of theory. It has been found, in all cases, that the direct association of Zn2+ with the carbonyl and thiocarbonyl groups takes place at the heteroatom attached to position 2 irrespective of its nature. This preference has been attributed to the resonance effects caused by the nearest heteroatoms (oxygen and nitrogen). The most stable complexes correspond to structures with Zn2+ bridging between the heteroatom at position 2 or 4 of the 4‐ or 2‐enol (or the 4‐ or 2‐enethiol) tautomer and the dehydrogenated ring nitrogen atom, N3. Zn2+ association has a clear catalytic effect on the tautomerization processes which connect the oxo–thione forms with the enol–enethiol tautomers. Hence, although the enol–enethiol tautomers of oxazolidine and its thio derivatives should not be observed in the gas phase, the corresponding Zn2+ complexes are the most stable species and should be accessible, because the tautomerization barriers are smaller than the Zn2+ binding energies. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
Zn/Zn5(OH)8Cl2·H2O flower-like nanostructures was electrodeposited on the coated Zn with poly (N-methyl pyrrole) in 0.1 M Zn (NO3)2 and 0.1 M KCl solution. The morphology and the structure of the Zn/Zn5(OH)8Cl2·H2O were characterized by Field Emission Scanning Electron Microscopy (FESEM), Fourier transform infrared (FT-IR) spectroscopy and X-ray diffraction analysis (XRD). The FT-IR results showed special peaks at 908 and 728 cm−1 related to Zn5(OH)8Cl2·H2O. The FESEM results indicated that Zn/Zn5(OH)8Cl2·H2O consists of a flower-like nanostructure and these flower-shaped structures contain many shaped nanopetals with the thickness of 27.8 nm. The XRD result confirmed that the major phase of electrodeposited product in 0.1 M KCl as supporting electrolyte was Zn5(OH)8Cl2·H2O. The ability of PMPy to create a thin film and the existence of several pores in its matrix act as a mold for the growth of Zn/Zn5(OH)8Cl2·H2O flower-like nanostructure. The trapping of Cl and OH within pores can be considered as the reason for the formation of flowerlike Zn/Zn5(OH)8Cl2·H2O nanostructures in 0.1 M KCl.  相似文献   

14.
采用M06-2X和CCSD(T)高阶量化计算和传统过渡态理论研究硫酸催化乙二醛气体相水化反应.对HCOCHO+H2O, HCOCHO+H2O+H2O, HCOCHO+H2O+H2O, HCOCHO+H2O...H2SO4和HCOCHO+H2O+H2SO4五个路径的反应机理和速率常数进行了研究.计算结果表明硫酸具有较强的催化能力,能显著减小乙二醛水化反应的能垒,在CCSD(T)/6-311++G(3df,3pd)//M06-2X/6-311++G(3df,3pd)理论水平,当硫酸分子参与乙二醛水化反应时,反应能垒从37.15 kcal/mol减少至7.08 kcal/mol.在室温条件下,硫酸催化乙二醛水化反应的反应速率1.34×10-11 cm3/(molecule.s),是等量水分子参与乙二醛水化反应的速率的1012倍,大于乙二醛与OH自由基反应的反应速率1.10×10-11 cm3/(molecule.s).这表明大气条件下,硫酸催化乙二醛水化反应可以发生,同乙二醛与OH自由基反应相竞争.  相似文献   

15.
The KLL and KLX Auger spectra of Cd from the EC-decay of 111In have been analyzed with the 7 eV instrumental resolution using a combined electrostatic spectrometer and a source prepared by a modified Langmuir–Blodgett method. Energies and relative intensities of all the nine well resolved KLL lines have been precisely determined. While the latter quantities were found to be in very good agreement with the relativistic intermediate coupling calculations, the former ones deviate from results of the widely used semi-empirical calculations especially in the case of the absolute energies. The predicted strong influence of the relativistic effects on the KL1L2(3P0) transition intensity has been proved. Most of the main KLX Auger lines have been well separated and their complex structures have been observed. Accurate energies and intensities obtained for these lines have been compared with results of calculations. The intermediate coupling splitting of the KL2L2 and KL3L3 lines have been revealed. The KLX/KLL, KLN(O)/KLM, and KXY/KLL group intensity ratios have also been determined. A conclusion is made that further finer theoretical investigation on the complex interaction mechanism for two inner-shell vacancies is still very much needed.  相似文献   

16.
Friedel–Crafts acylation and alkylation reactions were investigated using density functional theory calculations. The reaction systems studied were (benzene + acetyl chloride + Al2Cl6 (or AlCl3)) and (benzene + 2‐chloropropane + Al2Cl6). In the acylation reaction, the acylium ion intermediate is reached either via a Me? C(Cl)?O? Al2Cl6 complex or via direct Cl transfer: Me? C(?O)Cl? Al2Cl6 → Me? C?O?+? Al2Cl. The ion adds to benzene electrophilically to form a Wheland intermediate containing a strong C? H? Cl hydrogen bond, which leads to deprotonation and the subsequent formation of acetophenone. The resulting H? Cl? Al2Cl6 fragment is subjected to a nucleophilic attack by the carbonyl oxygen of the acetophenone, and recovery of the Al2Cl6 bridge is unlikely. Attack of the Al2Cl6 moiety by Me? C(Cl)?O gives the complex Me? C(Cl)?O–AlCl3, whose reactivity toward acylation is similar to that of the Me? C(Cl)?O–Al2Cl6 complex. In the alkylation reaction, deprotonation does not take place, but rather a [1,2] H‐shift from the Wheland intermediate. The resulting α‐protonated cumene undergoes deprotonation, with subsequent recovery of the Al2Cl6 bridge. In addition, the Al2Cl6‐catalyzed isomerization of the n‐propyl to the isopropyl cation was found to be a dyotropic shift. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
The reaction of PhCOCH2Cl with OH gave the expected α‐substituted alcohol (PhCOCH2OH) in addition to three dimer products. To clarify whether the substitution product is formed by direct SN2 or via carbonyl addition, the reaction of PhCOCH2Cl and OMe was examined. The reaction gave two products, PhCOCH2OH as the major product after acid hydrolysis and PhCOCH2OMe as the minor product. An electron‐withdrawing substituent on the phenyl ring enhanced the overall reactivity and gave more alcohol than ether. It was concluded that the alcohol was formed via carbonyl addition‐epoxidation route, whereas the ether was formed by the direct substitution route. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
Patrick Meier 《Molecular physics》2013,111(23):3859-3873
Three different approaches for calculating Franck–Condon factors beyond the harmonic approximation are compared and discussed in detail. Duschinsky effects are accounted for either by a rotation of the initial or final wavefunctions – which are obtained from state-specific configuration-selective vibrational configuration interaction calculations – or by a rotation of the underlying multi-dimensional potential energy surfaces being determined from explicitly correlated coupled-cluster approaches. An analysis of the Duschinsky effects in dependence on the rotational angles and the anisotropy of the wavefunction is provided. Benchmark calculations for the photoelectron spectra of ClO2, HS?2 and ZnOH? are presented. An application of the favoured approach for calculating Franck–Condon factors to the oxidation of Zn(H2O)+ and Zn2(H2O)+ demonstrates its applicability to systems with more than three atoms.  相似文献   

19.
ZnO nanoparticles, of average size of 10–15 nm, homogeneously dispersed in a silica matrix were prepared by a two stage citric acid/sol–gel process and thermal treatments up to 700 °C. The precursors formed at the early stages of the synthesis and their thermal evolution were investigated by FTIR, 29Si NMR MAS and CPMAS, 13C CPMAS and T(1H). A unidentate complex was revealed in the gel, together with other complexes in which citrate carboxylate groups are bound to one Zn2+ ion or act as a bridge between two Zn2+ ions. A comparison of the results from nanocomposite and silica samples prepared by the same method showed that chemical interactions between amorphous silica and zinc ions are not present either in the precursors or in the final materials. As a consequence, ZnO particles do not react with silica matrix when they are heated up to 700 °C. This result is ascribed to a nanophase segregation of zinc citrate complexes from the host matrix.  相似文献   

20.
Hydrogen bonding between teh protein and one or both of the two 1,4-quinone carbonyl groups of a benzo-or naphtho-quinone constitutes a significant protein-cofactor interaction in photosynthetic reaction centers. The redistribution of charge and spin density due to a particular H-bonding scheme leaves the largest hyperfine couplings (hfc) at the highest density positions, i.e., the nuclei of the carbonyl groups directly involved in H-bonding. The spin density changes at the ring carbon positions are accessed exeripmentlaly via electron paramagnetic resonance-determined hfc tensor elements of selective13C isotope labels in one of the two carbonyl groups. Complete hfc tensor data are presented for each of the13C positions in the functional charge-separated state in reaction centers of phytosystem I (PS I) isolated from cyanobacteria. A highly asymmetric H-bonding scheme for the A1 quinone binding site due to a single dominant H-bond to one carbonyl group is confirmed. A comparison to other wel-studied quinone binding sites of other protien-cofactor systems with more complex H-bonding schemes reveals the uniqueness of the PS I site. The single-sided A1 quinone site provides an ideal test case for the various sets of density functional theory (DFT) calculations that are currently available. While the overall agreement between experimental and calculated data is quite satisfactory, a significant discrepancy is found for the high-spin-density13C position associated with the H-bonded carbonyl. The dominant hfc component (and spin density) is underestimated in the DFT calculations, not only for the high-asymmetry case in PS I, but also for other quinone binding sites with less asymmetry that result from more complex H-bonding schemes. The cosnequences and potential relevance of this finding for biological function are discussed.  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