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 three azaborine isomers with the formula C4H6BN, 1,2-, 1,4-, and 1,3-azaborine ( I , II , and III ), have been examined using MNDO (m odified n eglect of d iatomic o verlap) calculations. The most stable azaborine was I (heat of formation -8.147 kcal/mol), followed by II (+11.60 kcal/mol) and III (+16.64 kcal/mol). Qualitatively, although the π- and π*-orbitals calculated for the azaborines exhibited an ordering similar to that in benzene and borazine, the HOMO/LUMO energy differences (9.27, 9.68, and 8.44 eV, respectively) were smaller than was the difference calculated for borazine (12.81 eV), but of the same magnitude as the difference for benzene (9.76 eV). With the exception of borazine, each molecule had a π-orbital for the HOMO and a π*-orbital for the LUMO ; borazine's LUMO was a π*-orbital. The calculated shapes and atomic contributions for the π-and π*-orbitals of the azaborines were best described as “hybrids” of the π- and π*-orbitals of benzene and borazine. As was observed for the π- and π*-orbitals of borazine, the azaborines exhibited increased orbital density at the nitrogen atom in the π-bonding orbitals and at boron in the π-antibonding orbitals, as would be predicted from electronegativity considerations. Although I and II exhibited significant double- and single-bond localization, all of the ring bonds in III were delocalized. The delocalization in III was not uniform but, rather, resembled two inequivalent fused allyl systems. The cyclotrimer ( IV ) of 1,2-azaborine (heat of formation -44.07 kcal/mol), based purely on thermodynamic considerations, was predicted to form spontaneously from three monomer molecules with the concurrent loss of three molecules of dihydrogen. The cyclotrimers that could theoretically be produced from 1,2-azaborine without the loss of dihydrogen ( IVc and IVt ) were each calculated to be less stable (heats of formation +24.45, and +33.29 kcal/mol, respectively) than was the experimentally observed IV . The carbon molecules triphenylene ( TP ) and cis- and trans-4a,4b,8a,8b,12a,12b- hexahydrotriphenylene ( TPc and TPt ) (heats of formation +76.79, +101.6, and +103.1 kcal/mol, respectively) were each calculated to be less stable than were the azaborine cyclotrimer analogs, as was observed in comparisons of benzene with the azaborines and borazine.  相似文献   

2.
The relative aromaticities of the three singlet benzyne isomers, 1,2-, 1,3-, and 1,4-didehydrobenzenes have been evaluated with a series of aromaticity indicators, including magnetic susceptibility anisotropies and exaltations, nucleus-independent chemical shifts (NICS), and aromatic stabilization energies (all evaluated at the DFT level), as well as valence-bond Pauling resonance energies. Most of the criteria point to the o-benzyne相似文献   

3.
The influence of the relative boron and nitrogen positions on aromaticity of the three isomeric 1,2‐, 1,3‐, and 1,4‐azaborines has been investigated by computing the extra cyclic resonance energy, NICS(0)πzz index and by visualizing the π‐electron (de)shielding pattern as a response of the π system to a perpendicular magnetic field. The origin of the known stability trend, in which the 1,2‐/1,3‐isomer is the most/least stable, was examined by using an isomerization energy decomposition analysis. The 1,3‐arrangement of B and N atoms creates a charge separation in the π‐electron system, which was found to be responsible for the lowest stability of 1,3‐azaborine. This charge separation can, in turn, be considered as a driving force for the strongest cyclic π‐electron delocalization, making this same isomer the most aromatic. Despite the well‐known fact that the B?N bond attenuates electron delocalization due to large electronegativity difference between the atoms, the 1,4‐B,N relationship reduces aromaticity to a greater extent by making the π‐electron delocalization more one‐directional (from N to B) than cyclic. Thus, 1,4‐azaborine was found to be the least aromatic. Its lower stability with respect to the 1,2‐isomer was explained by the larger exchange repulsion.  相似文献   

4.
Rotational isomers (rotamers) of the three structural isomers of benzenedicarboxaldehydes (1,2-, 1,3-, and 1,4-derivatives) have been investigated in detail using matrix-isolation infrared spectroscopy in the 600-4000 cm-1 region, combined with UV photoexcitation and density-functional theory (DFT) calculations. Two rotamers were identified for 1,2- and 1,4-benzenedicarboxaldehyde (1,2- and 1,4-BDA, respectively), while three rotamers were identified for 1,3-benzenedicarboxaldehyde (1,3-BDA) in infrared spectra upon UV-irradiation. Most of the observed infrared bands of each rotamer have been assigned. The energetic relationships among the rotamers were revealed based on the infrared data and the DFT calculations. It is shown that the intramolecular C-H...H-C interaction in the H-syn rotamer or the C-H...O=C hydrogen bonding in the anti rotamer of 1,2-BDA results in the blue-shift of the aldehyde C-H stretching band and the shortening of the aldehyde C-H bond length. Both photoinduced rotational isomerization and rearrangement were observed upon UV irradiation for 1,2-BDA. The structure of the major enol isomer formed as the result of the photochemical rearrangement of 1,2-BDA is determined.  相似文献   

5.
6.
采用动态气体提取方法, 使用自行研制的质子转移反应质谱(PTR-MS)装置, 研究了挥发性苯系物亨利常数的测量方法. 考察了苯在提取气体与水溶液中达到平衡需要的液面高度, 通过25 ℃条件下苯的亨利常数测量验证了动态气体提取方法结合质子转移反应质谱测量亨利常数的有效性, 报告了11种苯系物的亨利常数测量结果. 与常规的亨利常数测量方法相比, 本方法具有简单、快速、人为干扰少等优点.  相似文献   

7.
The molecular structures, total energies, and other computational data of benzene, and its monoand dihalogenated derivatives (halogen = F, Cl, Br) have been studied by DFT calculations. The main aim of the study was to estimate the relative stabilities (energies) of the ortho, meta, and para isomers of the six series of dihalobenzenes investigated. The computational data show that the ortho isomers always have the highest, and the meta isomers usually, but not always, the lowest total energies. Thus, 1,2-difluorobenzene is ca. 16.6 kJ mol–1, and 1,4-difluorobenzene 2.5 kJ mol–1 less stable than 1,3-difluorobenzene. Among the other isomeric dihalobenzenes, the differences in stability are less pronounced. For the dibromo-, dichloro-, and bromochlorobenzenes, the para compounds are calculated to be slightly (0.2–0.4 kJ mol–1) more stable than their meta isomers. In addition to the thermochemical aspect of the study, the computational molecular structures of the halobenzenes are compared with available experimental data and discussed in terms of the substituent-induced deformation of the ideal geometry of the benzene ring. The computational electric dipole moments, especially for the fluorine-containing compounds, compare favorably with the respective experimental (gas-phase) values.  相似文献   

8.
The isomerization of tetraaryl tetrakis(trifluoromethyl)[4]radialenes was studied. When type II (all-Z) isomers of 5,6,7,8-tetraaryl-5,6,7,8-tetrakis(trifluoromethyl)[4]radialenes were heated in tetralin at 170-200 degrees C, isomerization occurred to give mixtures of four [4]radialenes in a ratio of ca. I:II:III:IV = 1:10:5:1. However, when the isomeric mixtures were heated in the solid state at the same temperature, selective isomerization took place to give type II isomers in good selectivity (>91%). Upon irradiation with light, the type II isomers first isomerized to mixtures of the four [4]radialene isomers (I:II:III:IV = 2:2:48:48) and then rearranged to cyclobuta[b]naphthalenes via a 6pi-electrocyclic reaction followed by 1,3-hydrogen migration.  相似文献   

9.
Diels–Alder reactions employing 1,2‐azaborine heterocycles as 1,3‐dienes are reported. Carbocyclic compounds with high stereochemical and functional complexity are produced, as exemplified by the straightforward two‐step synthesis of an amino allyl boronic ester bearing four contiguous stereocenters as a single diastereomer. Whereas electron‐deficient dienophiles undergo irreversible Diels–Alder reactions, a reversible Diels–Alder reaction with the less electron‐deficient methyl acrylate is observed. Both the N and the B substituent of the 1,2‐azaborine exert significant influence on the [4+2] cycloaddition reactivity as well as the aromatic character of the heterocycle. The experimentally determined thermodynamic parameters of the reversible Diels–Alder reaction between 1,2‐azaborines and methyl acrylate correlate with aromaticity trends and place 1,2‐azaborines approximately between furan and thiophene on the aromaticity scale.  相似文献   

10.
Ab initio calculations using 6-311G**, cc-pVDZ, aug-cc-pVDZ, and a (valence) double-zeta pseudopotential (DZP) basis set, with (MP2, QCISD, CCSD(T)) and without (UHF) the inclusion of electron correlation, and density functional methods (B3LYP) predict that 1,n-homolytic transfers (n = 1-5) of silyl, germyl, and stannyl groups from group IV heteroatoms to carbon radicals can proceed via a frontside attack mechanism. At the B3LYP/DZP level of theory, energy barriers (DeltaE++) of 101.2, 98.8, 58.9, and 63.4 kJ/mol are calculated for the 1,2-, 1,3-, 1,4-, and 1,5-translocation reactions, respectively, of SiH3 between silicon atoms. Similar results are obtained for reactions involving germanium and tin with energy barriers (DeltaE++) of 85.9-113.1, 84.4-109.0, 41.7-73.3, and 48.5-78.2 kJ/mol for the 1,2-, 1,3-, 1,4-, and 1,5-translocation reactions, respectively. This study also predicts that four- and five-membered ring-closure reactions can be competitive with the 1,4- and 1,5-translocation reactions. These results suggest that while 1,2- and 1,3-translocation four-membered ring-formation reactions are unlikely to be synthetically viable, 1,4- and 1,5-transfers and five-membered ring-formation have synthetic possibilities.  相似文献   

11.
Co(cycloheptylamine)2Ni(CN)4 host complex has been prepared in powder form and its vibrational spectra are investigated. The vibrational assignment of the cycloheptylamine in various phases (gas, solution, liquid and complex) are facilitated by means of DFT calculations performed on the free ligand molecule. The spectral data suggest that the prepared complex is similar in structure to the Hofmann-dma-type hosts. The sorption processes of some aromatic guests (benzene, toluene, o-, m-, p-xylene, 1,2-, 1,3-dichlorobenzene, 1,4-dichlorobenzene, 1,4-dibromobenzene and naphthalene) in this host complex have also been examined at room temperature by gravimetric and spectroscopic measurements. The desorption of the benzene guest against time has been measured. The structure of the Co(cycloheptylamine)2Ni(CN)4 host changes on inclusion of the guest molecule and recovers after liberation.  相似文献   

12.
    
Co(cycloheptylamine)2Ni(CN)4 host complex has been prepared in powder form and its vibrational spectra are investigated. The vibrational assignment of the cycloheptylamine in various phases (gas, solution, liquid and complex) are facilitated by means of DFT calculations performed on the free ligand molecule. The spectral data suggest that the prepared complex is similar in structure to the Hofmann-dma-type hosts. The sorption processes of some aromatic guests (benzene, toluene, o-, m-, p-xylene, 1,2-, 1,3-dichlorobenzene, 1,4-dichlorobenzene, 1,4-dibromobenzene and naphthalene) in this host complex have also been examined at room temperature by gravimetric and spectroscopic measurements. The desorption of the benzene guest against time has been measured. The structure of the Co(cycloheptylamine)2Ni(CN)4 host changes on inclusion of the guest molecule and recovers after liberation.  相似文献   

13.
The distribution of isomers at the preparation of substituted aryldichlorophosphines using AlCl3 was determined by means of IR spectroscopy. The isomers of the methylphenyl- and chlorophenyl dichlorophosphines as well as the 1,4-fluorophenyl dichlorphosphine (prepared for the first time) were prepared by way of the corresponding substituted phenylphosphonic bis(diethylamides). Thus calibration curves were obtained for quantitative evaluation. Alkyl- as well as halogene benzenes yield mainly 1,4-substituted aryl dichlorophosphines. Halogenobenzenes yield little 1,2 and little 1,3 isomers, whereas up to 40% 1,3 isomers with little 1,2 product are obtained from alkylbenzenes. *** DIRECT SUPPORT *** A3615112 00022  相似文献   

14.
The previously reported uranium(III) complex [(((Ad)ArO)(3)N)U(III)(DME)] (1; Ad = adamantane, DME = 1,2-dimethoxyethane) reacts with the terminal bis-alkynes 1,7-octadiyne or 1,6-heptadiyne in C-C-coupling reactions to form the uranium(IV) vinyl complexes [{(((Ad)ArO)(3)N)U(IV)}(2)(μ-η(2):η(1)-1,2-(CH)(2)-cyclohexane)] (2) and [{(((Ad)ArO)(3)N)U(IV)}(2)(μ-η(2):η(2)-1,2-(CH)(2)-cyclopentane)] (3). With the monoalkynes 1-hexyne or 4-(t)butyl-phenylacetylene, the complexes [{(((Ad)ArO)(3)N)U(IV)}(2)(μ-η(2)(C1):η(1)(C4)-2-(n)Bu-1,3-octadiene)] (4) and [{(((Ad)ArO)(3)N)U(IV)}(2)(μ-η(2)(C4):η(1)(C1)-1,3-di-(p-(t)Bu-phenyl)butadiene))] (5), are formed. These are the first four examples of uranium vinyl complexes that are reported and crystallographically characterized. In addition, detailed DFT calculations are presented to establish a possible mechanism for their formation and explain the differences found for the coordination of the hydrocarbon fragments. In contrast to a previously proposed monometallic pathway for catalytic hydroamination of alkynes and alkyne dimerization involving a uranium vinyl intermediate at uranium(III) complexes, the calculations clearly support a bimetallic mechanism, since its transition states are energetically the most favored.  相似文献   

15.
3,4,5,6-Tetrafluoro-2-nitroaniline (I), 2,3,5,6-tetrafluoro-4-nitroaniline (IV) and 2,5,6-trifluoro-4-nitro-1,3-phenylenediamine (VI) react with nitrous acid to give 3,4,5-trifluoro-6-nitro-1,2-diazo-oxide (III), 3,5,6-trifluoro-4-nitro-1,2-diazo-oxide and (V) 5-fluoro-6-nitro-bis-1,2:3,4-diazo-oxide (VII), respectively. Reduction of the diazo-oxide (III) with hypophosphorous acid gives 4,5,6-trifluoro-3-nitrophenol (VIII). Treatment of 2,3,4,6-tetrafluoroacetanilide with nitric acid affords trifluoro-p-benzoquinone (X), the reduction of which gives trifluorohydroquinone (XI). Proton and fluorine chemical shifts and coupling constants of the new compounds are reported.  相似文献   

16.
A colorimetric sensing system based on gold nanoparticles functionalized with a water‐soluble anion‐recognition motif, an o‐(carboxamido)trifluoroacetophenone analogue, has been developed. The nanoparticle system selectively senses specific isomers of dicarboxylates that are geometrically favorable for the binding‐induced aggregation process; thus, it discriminates a trans‐dicarboxylate fumarate from its cis‐isomer maleate, and benzene‐1,4‐dicarboxylate from its isomeric benzene‐1,2‐and benzene‐1,3‐dicarboxylates in water, exhibiting a color change from red to blue.  相似文献   

17.
采用DFT,QCISD和CCSD(T)等理论计算方法对三重态SiCP2异构体的结构和稳定性进行了理论研究.在B3LYP/6-311G(d)水平下,共计算得到由17个过渡态相连接的15个异构体.在CCSD(T)/6-311 +G(2df)//QCISD/6-311G(d)水平下,考虑重点振动能相对能量最低的三元环状异构体P-cCSiP 8(0.0 kJ/mol)及四元环状结构的cPCSiP 4具有相当大的动力学稳定性,在一定的实验室和星际条件下可能被检测到.另外,对它们的成键性质也进行了分析.  相似文献   

18.
Cycloaddition of sulfene to N,N-disubstituted 4-amino-3-phenyl-3-buten-2-ones (III) occurred in good yield only in the case of aliphatic N-substitution to give 4-dialkylamino-3,4-dihydro-6-methyl-5-phenyl-1,2-oxathiin 2,2-dioxides, whereas N,N-disubstituted 4-amino-1-phenyl-3-buten-2-ones (IV) did not react at all. Polar 1,4-cycloaddition of dichloroketene to III and IV occurred partly in the case of aromatic N-substitution, with the exception of the morpholino derivative IVd, giving in low yield N,N-disubstituted 4-amino-3,3-dichloro-3,4-dihydro-(6-methyl-5-phenyl)(6-benzyl)-2H-pyran-2-ones, which were dehydrochlorinated with DBN to the corresponding 4-amino-3-chloro-(6-methyl-5-phenyl)(6-benzyl)-2H-pyran-2-ones (VII) in good yield. In some cases of aliphatic N,N-disubstitution of III and IV, cycloaddition led directly to N,N-dialkyl derivatives VII in low yield.  相似文献   

19.
The Diels-Alder reaction of fluoranil with cyclopentadiene, 1,3-butadiene, and 1-acetoxy-1,3-butadiene gave 1,4, 5, 8-bis(methylene)-4a, 8a, 9a, 10a-tetrafluoro-1, 4, 4a, 5, 8, 8a, 9a, 10a-octahydroanthraquinone (I), 2, 3, 4a, 8a-tetrafluoro-4a, 5, 8, 8a-tetrahydro-1,4-naphthoquinone (III), and 5-acetoxy-2, 3, 4a, 8a-tetrafluoro-4a, 5, 8, 8a-tetrahydro-1,4- naphthoquinone (VI), respectively. Hydrogenation of I gave the expected saturated diketone(II). Hydrogenation of III afforded, with elimination of the two tertiary fluorines, 2,3-difluoro-5, 6, 7, 8-tetrahydro-1, 4- dihydroxynaphthalene (IV). In hydrogenation of VI, acetic acid and two moles of hydrogen fluoride were eliminated to give 2,3-difluoro-1, 4-dihydroxynaphthalene(VII). Both dihydroxy compounds IV and VII yielded on oxidation with ferric chloride the corresponding quinones, 2, 3- difluoro-5, 6, 7, 8-tetrahydro-1, 4-naphthoquinone (V) and 2, 3-difluoro-1, 4-naphthoquinone (VIII), respectively. Equivalent amounts of compounds IV and V gave a red-brown semiquinone IX, and a mixture of VI and VIII gave a dark-violet semiquinone X.  相似文献   

20.
Novel blue emitters were synthesized based on the fullerene fragment corannulene. 1,2- bis(corannulenylethynyl)benzene and 1,4-bis(corannulenylethynyl)benzene were designed, synthesized, and shown to exhibit significant red shifts in their absorption spectra as compared to that of the parent corannulene. Photoluminescence studies show both 1,2- bis(corannulenylethynyl)benzene and 1,4- bis(corannulenylethynyl)benzene gives enhanced blue luminescence compared to the parent corannulene structure. 1,4-bis(corannulenylethynyl)benzene was observed to give intense blue luminescence when excited at 400 nm. DFT and TD-DFT calculations were performed and shown to be consistent with the observed experimental results.  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