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 259 毫秒
1.
Methylenecyclopropane 1 undergoes photochemical rearrangement to 2. Intervention of cyclobutylidene 3 explains not only the rearrangement but also newly obtained products such as cyclobutene 4 and cyclobutylidenecyclobutane 6. Experiments designed to generate cyclobutylidene 3 independently have provided some support for the intermediacy of 3.  相似文献   

2.
Photochemistry and thermal reaction of 2,2-diphenylmethylenecyclopropane 1a have been reinvestigated to make their mechanistic refinement. A new finding in the photochemistry of 1a is the detection and isolation of cyclobutene 7. While fragmentation and 1,3-carbon (C) shift is responsible for previously reported photoproducts, 1,1-diphenylethylene 3 and diphenylmethylenecyclopropane 2a, respectively, a new pathway is required to explain the formation of cyclobutene 7. A mechanism involving the 1,2-C shift (Scheme 3, arrow b) followed by 1,2-hydrogen (H) shift is proposed. In the thermal reaction of 1a, on the other hand, a trimethylenemethane type of species is shown to be a common intermediate for degeneracy, rearrangement from 1a to 2a, and formation of indene 5. In the photochemistry of 1a, intervention of cyclobutylidene 8 is strongly supported by circumstantial evidences provided by steady-state and laser flash photolytic investigations. Further experiments designed to independently generate cyclobutylidene 8 showed the methylenecyclopropane/cyclobutene branching ratio (1a/7) to be in the range between 0 and 1.7, which is much lower than the value of 4.0 for parent cyclobutylidene. Nevertheless, the relative efficiency of 1,2-shift pathway to generate 8 is shown to be as high as 17–47% compared to 1,3-C shift to give 2a at the early photochemical stage of 1a.  相似文献   

3.
《Polyhedron》1999,18(6):905-908
The aluminum alkoxides [MeClAlOEt]3 (1), [Et2AlOMe]3 (2), [Me2AlOEt]3 (3), and [EtClAlOEt]3 (4) were investigated by 1H NMR spectroscopy to study the o-dichlorobenzene solution equilibrium: 2 [R2AlOR′]3⇌3 [R2AlOR′]2. The complexes are shown to exist primarily as trimers at room temperature, but increasing concentrations of the dimeric form are observed at higher temperatures. Equilibrium constants, ΔH, and ΔS were determined for the trimer–dimer equilibrium. Values of ΔH for the conversion of 2 moles of trimer are 63(4), 78(1), 85.0(8), and 99(6) kJ for 1, 2, 3, and 4, respectively. The corresponding values of ΔS are 142(7), 184(3), 218(2), and 265(17) J/K, respectively. Thermodynamic parameters are compared with those reported for [Me2AlOPrn]3 and [Me2AlOPh]3. The characterization of [EtClAlOMe]3 is also reported.  相似文献   

4.
The reaction of [Os3(CO)12 with [12]aneS3 ([12]aneS3  {(CH2)3S}3) in octane for 6 h, under reflux, led to isolation of two products [Os3(CO)11([12]aneS3)] (1) and [Os4(CO) 13([12]aneS3)] (2), while with [Ru3(CO)12] under similar conditions, in THF, a number of products were obtained, including [Ru4(CO)11([12]aneS3)] (3), [Ru5(CO)13([12]aneS3)] (4), and [Ru6(CO)16([12]aneS3)] (5). An X-ray diffraction study of 2 shows that the macrocycle is coordinated to the ‘wingtips’ of an Os4 butterfly through the two electron pairs on one sulphur atom, while in 5 all three sulphur atoms of the macrocycle coordinate to two of the Ru atoms in a spiked edge-bridged tetrahedral metal framework.  相似文献   

5.
L. Lombardo  D. Wege 《Tetrahedron》1974,30(21):3945-3952
The title reaction gave a 2+2 cycloadduct, 8,9-benzo-cis-bicyclo[5.2.0]nona-2,4,8-triene 7, together with ene product, 7-phenylcycloheptatriene. The structure of 7 was confirmed by catalytic reduction to give 8,9-benzo-cis-bicyclo[5.2.0]non-8-ene, which was also obtained in the reaction of benzyne with cycloheptene, and by reduction of the known 8,9-benzobicyclo[5.2.0]nona-1,8-diene. Other benzo(C9H10) hydrocarbons which have been synthesised are 7,8-benzobicyclo[4.2.1]nona-2,4,7-triene 5, 2,3-benzobicyclo[6.1.0]nona-2,4,6-triene 28 and 4,5-benzobicyclo[6.1.0]nona-2,4,6-triene 29. The thermolysis of 7, 28, 29 and of 3,4-benzo-exo-endo-tetracyclo[4.3.1.03,4.07,9]dec-3-en-10-one, 25, is described.  相似文献   

6.
Four oxovanadium and one dioxovanadium complex with 2-hydroxyacetophenone N(4)-phenylthiosemicarbazone (H2L) which are represented as [VOLphen]·2H2O (1), [VOLbipy] (2), [VOLdmbipy] (3), [VOL]2 (4) and [VO2HL]·CH3OH (5) have been synthesized and characterized by elemental analyses, electronic, infrared and EPR spectral techniques. In all the complexes 14 the ligand coordinates through phenolic oxygen, azomethine nitrogen and thiolate sulfur. But in complex [VO2HL]·CH3OH, coordination takes place in thione form instead of thiolate sulfur. All the complexes except [VO2HL]·CH3OH are EPR active due to the presence of an unpaired electron. In frozen DMF at 77 K, all the oxovanadium(IV) complexes show axial anisotropy with two sets of eight line patterns.  相似文献   

7.
The disodium salt of a bis(aryloxide)-N-heterocyclic-carbene dianionic ligand, Na2[L], was prepared by reaction of 1,3-bis(4,6-di-tert-butyl-2-hydroxybenzyl)imidazolium bromide [H3L]Br with 3 equiv. of NaN(SiMe3)2. Reaction of ZrCl4(thf)2 with 1 equiv. of Na2[L] gave a mixture of [L]ZrCl2(thf) (1) and [L]2Zr (2). When the amount of Na2[L] was increased to 2 equiv., the bis(bis(aryloxide)-N-heterocyclic-carbene) complex 2 was obtained in good yield. The dichloro complex 1 is a precursor to organometallic derivatives, and treatment with PhCH2MgCl or Me3SiCH2Li yielded [L]ZrR2 [R = CH2Ph (3), CH2SiMe3 (4)]. The disodium salt of the ligand Na2[L] is unstable and undergoes 1,2-benzyl migration, whereas zirconium complexes of the [L]2− ligand are found to be thermally stable in solid and solution. The X-ray crystal structures of 1, 2, and 3 are described.  相似文献   

8.
Reaction of titanium cyclobutylidene complexes, prepared by the desulfurizative titanation of 1,1-bis(phenylthio)cyclobutanes with Cp2Ti[P(OEt)3]2, with alkynes gave 1-(alk-1-enyl)cyclobutenes.  相似文献   

9.
A.T. Bottini  L.J. Cabral 《Tetrahedron》1978,34(21):3195-3199
Dispiro[2.0.2.2]oct-7-ene 1 was synthesized by debrominatioa of cis- and trans-7,8-dibromodispiro[2.0.2.2]octane 3a with LAH and by dechlorination of cis- and trans-7,8-dichlorodispiro[2.0.2.2]octane 3b with magnesium. Stepwise electrophilic additions to 1 of HBr, HI, Br2 and Cl2 were studied. The major products (and yields) from these reactions were: 7-bromodispiro[2.0.2.2]octane 2a (43%), 4-iodo-4,5-ethanospiro[2,3]hexane 4b (ca. 50%); trans-3a (40%); and cis-3b (20%). Free-radical addition of hydrogen bromide to 1 gave an 80% yield of 7-bromodispiro[2.0.2.2]octane 2a. At ?10°, hydroboration-oxidation of 1 was found to give mainly 7-hydroxydispiro[2.0.2.2]octane 2a in ca. 90% yield; at 25°, near equal amounts of 2c and 4-(2-hydroxyethyl) spiro[2.3]hex-4-ene 14 were obtained.  相似文献   

10.
Metalation of (RSiMe2)3CH (1a R = H, 1b R = Me, 1c R = Ph) with lithium diisopropylamide (LDA) or methyllithium in THF gave organolithium reagents (RSiMe2)3CLi, which reacted with the formylated calixarene (2), to give the corresponding 5,17-bis[2,2-bis(organosilyl)-1-ethenyl]-25,26,27,28-tetrapropoxycalix[4]arenes (3a, 3b and 3c) via the Peterson olefination. The compounds (RSiMe2)3CLi were treated with 25,26,27,28-tetrakis(4-bromobutoxy)calix[4]arene (4) to give 25,26,27,28-tetrakis[4-(tris(dimethylsilyl)methyl)butoxy] calix[4]arene (5a) and 25,26,27,28-tetrakis[4-(tris(trimethylsilyl)methyl)butoxy] calix[4]arene (5b) via nucleophilic substitution reactions. However the compound 25,26,27,28-tetrakis[4-(tris(dimethylphenylsilyl)methyl)butoxy] calix[4]arene (5c) was not obtained, presumably because (PhSiMe2)3C- is highly sterically hindered and the reactivity of its derivatives is low. The compound 5a has potential as a core for dendrimers.  相似文献   

11.
2,15 - Dimethoxycarbonyl - [15]annulenone 4,7:10,13 - dioxide (14) has been prepared by the condensation of cis -α,β -bis(5 - formyl - 2 - furyl)ethy]ene (13) with dimethyl acetonedicarboxylate. Treatment of 14 with conc H2SO4 led to 4,7:10,13-dioxido[15]annulenone 2,15-dicarboxylic acid anhydride (17), which was subsequently converted to the corresponding dicarboxylic acid (18) by dilute KOH. Decarboxylation of 18 gave rise to two isomeric [15]annulenone 4,7:10,13 - dioxides, i.e., the tri-cis isomer (7) and the mono - trans - di - cis isomer (8).Regarding to the ring current effects, the proton chemical shifts of these [15]annulenones were compared with those of a reference model, 4,7:10,13 - dioxido - cyclopentadecaheptaene (2.4.6.8.10.12.14) (3). Both of the parent [15]annulenones (7 and 8) have been interpreted as nondiatropic, while the anhydride (17) has been shown to be diatropic, sustaining an induced diamagnetic ring current. The enforced planarity and symmetrical geometry of the anhydride have been discussed. As expected, when 7, 8, 14, 17 and 18 were dissolved in CF3COOH or conc H2SO4, completely delocalized [15]annulenium cations were produced, all of which proved to be diatropic. Three possible geometrical isomers of these 14π cations were established experimentally.  相似文献   

12.
The reduction of Mn2(CO)7(μ-S2), (1) with sodium amalgam in THF provided the new monoanion [Mn3(CO)103-S2)2], (3) isolated in low yield as the [Ph3PMe] salt. The reaction of Mn4(CO)153-S2)(μ4-S2), (2) with [Ph3PMe]I provided the same salt [Ph3PMe] [3] in a good yield, 68%. Anion 3 reacts [CpFe(CO)2(acetone)]BF4 to yield the neutral mixed metal complex CpFeMn3(CO)123-S2)(μ4-S2), (4). The structures of [Ph3PMe] [3] and 4 were determined by single crystal X-ray diffraction analyses. The core of the structure of 3 consists of two [Mn(CO)3] groups bridged by two disulfido ligands in a μ2–η2 fashion with an additional [Mn(CO)4] group that bridges the two disulfido ligands. The CpFe(CO)2 group in 4 is bonded to a sulfur atom of one of the two disulfido ligands of the anionic grouping of 3.  相似文献   

13.
Several novel azacalix[4]aromatics constituting terphenylene units have been synthesized via sequential nucleophilic aromatic substitution reactions of 5′-t-butyl-(1,1′:3′,1″-terphenyl)-3,3″-diamine 9 and 5′-t-butyl-(1,1′:3′,1″-terphenyl)– 4,4″-diamine 11 with 1,5-difluoro-2,4-dinitrobenzene and cyanuric chloride, respectively. The bridging –NH– functions of the tetra-nitro substituted azacalix[2]arene[2]terphenylenes 1 and 2 have been transformed to the corresponding –N(CH3)– bridged azacalix[2]arene[2]terphenylenes 3 and 4 via N-alkylation. Single crystal X-ray analysis revealed that the terphenyl-3,3″-diamine derived azacalix[2]terphenylene[2]triazine 5 adopts a distorted chair conformation in the solid state, and the terphenyl-4,4″-diamine derived azacalix[2]terphenylene[2]triazine 6 was found to adopt a 1,3-alternate conformation.  相似文献   

14.
1,6-Diazabicyclo[4.3.3]dodecane (7), 1,6-diazabicyclo[4.3.2]undecane (8) and 1,8-diazabicyclo[6.3.3]tridecane (9) have been made by reduction of tricyclic α-aminoammonium salts with LiAlH4; the protonation and oxidation of 8 and 9 are discussed.  相似文献   

15.
The layer-by-layer (LbL) assembled thin films containing tetraamino-thiacalix[4]arenes (1) and tetraamino-calix[4]arenes (2) were used as nanoreactor to synthesize in situ Ag nanoparticles (Ag NPs). UV–vis spectra and AFM images demonstrate that Ag NPs are included in the (1/Ag NPs)n and (2/Ag NPs)n multilayer films. The silver ions are absorbed through cation–π interaction and calix[4]arene-metal ion coordination interaction and are reduced into Ag NPs by calix[4]arenes. TEM images indicated that Ag NPs within aminocalix[4]arene multilayers were highly dispersed and uniform. Moreover, the mean size of Ag NPs is smaller than 10 nm.  相似文献   

16.
The intramolecular [2+2] cycloaddition of 1,3-dienes under visible light irradiation investigated by Yoon and his co-workers shows remarkably high yield and stereoselective differences under different photocatalysts. The reaction was speculated to be induced by energy transfer. However, the origin for these phenomena is still unclear. In this scene, the detailed mechanism for the [2+2] cycloaddition of 1,3-dienes under visible light has been investigated using density functional theory B3LYP and TPSSTPSS methods. The result shows that the reaction not only can be induced by energy transfer between photocatalysts and reactants, but also can be induced by electron transfer between them. The [2+2] cycloaddition induced by energy transfer is carried out along the potential energy surface (PES) of triplet excited states (T1) firstly, and then goes back to the singlet ground state (S0) via MECPs (minimum energy crossing points) between the PESs of the S0 and T1 states, forming the product in the S0 state. The [2+2] reaction induced by electron transfer proceeds along the doublet state PES of the cation radical reactant and the neutral four-membered ring product could be obtained by electron transfer from the corresponding reactant or reduced photocatalyst. The origin of stereoselectivity of the [2+2] reaction is attributed to the reaction mechanism difference under different photocatalysts.  相似文献   

17.
The reductive coupling reaction of 1,4-bis(3-acetyl-5-tert-butyl-2-methoxyphenyl)butane 3 was carried out using TiCl4-Zn in pyridine followed by a McMurry coupling reaction to afford the compounds anti and syn 1,2-dimethyl[2.4]MCP-1-ene 4. Bromination of 4 with BTMA-Br3 in dry CH2Cl2 afforded the interesting compound 1,2-bis-(bromomethyl)-5,15-di-tert-butyl-8,18-dimethoxy[2.4]MCP-1-ene 6 and consecutive debromination with Zn and AcOH in CH2Cl2 solution afforded the stable solid 5,15-di-tert-butyl-8,18-dimethoxy-1,2-dimethylene[2.4]MCP 7 in 89% yield. Compound 7 was conveniently employed in a Diels–Alder reaction with dimethyl acetylenedicarboxylate (DMAD) to provide 2-(3′,6′-dihydrobenzo)-5,15-di-tert-butyl-8,18-dimethoxy[2.4]MCP-4′,5′-dimethylcarboxylate 8 in good yield. Diels–Alder adduct 8 was converted into a novel and inherently chiral areno-bridged compound [2.4]MCP 9 by aromatization. The chirality of the two conformers was characterized by circular dichroism (CD) spectra of the separated enantiomer which are perfect mirror images of each other.  相似文献   

18.
Here we report the design and syntheses of two new triptycene-based rigid acyclic C-shaped hosts, clip[5]arenes C[5]OH and C[5]ME, and the strong host–guest complexation between C[5]OH and an electron-poor bipyridinium salt, paraquat G. The Ka value for the host–guest complex C[5]OH???G was calculated to be (1.09?±?0.36)??×??105?M?1 in acetone by using a non-linear curve-fitting method based on the UV–vis absorption titration experiments. Furthermore, based on this new host–guest recognition motif, a novel pseudopolyrotaxane-like supramolecular structure was constructed with C[5]OH threaded on polyviologen polymer VP-10.  相似文献   

19.
Treatment of [Cp(CO)3Mo] with ClCOCF2COCl gave [Cp(CO)3MoCOCF2COMo(CO)3Cp] (1), which, by loss of CO gave [Cp(CO)3MoCOCF2Mo(CO)3Cp] (2). Further loss of CO moieties from 2 resulted in a mixture of different molybdenum complexes. The trinuclear complex [Cp3(CO)6Mo33-CF)] (3) was considered the most interesting one. The crystal structure of 3 is presented. The compound appears as two crystallographical independent molecules of identical structure, which crystallize in two enantiomeric forms.  相似文献   

20.
Treatment of the ruthenium complex [Ru]---

(3, [Ru]=Cp(dppe)Ru) containing a heterocyclic [1,3]-thiazine-4-thione six-membered-ring ligand with various organic halides results in alkylation at the thione sulfur terminus of the ligand to yield [Ru]---

][X] (4a, R=CN, X=I; 4b, R=Ph, X=Br; 4c, R=CH=CH2, X=I, 4d, R=p-C6H4CF3, X=Br). Similarly the reaction of 3 with HgCl2 at room temperature affords [Ru]---

][Cl] (5). Transformation of 5 to the cationic vinylidene complex {[Ru]=C=C(Ph)C(O)NHPh}2[Hg2Cl6] (6) readily occurred in the air. The structures of 4c and 6 are determined by single crystal X-ray diffraction analysis.  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