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.
A hexanuclear heterometallic cluster of composition [Dy2Co4(L)4(NO3)2(OH)4(C2H5OH)2] ⋅ 2 C2H5OH ( 1 ) was synthesized by employing a Schiff base 2-(((2-hydroxy-3-methoxybenzyl) imino)methyl)-4-methoxyphenol (H2L) as ligand and utilizing Dy(NO3)3 ⋅ 6H2O and Co(NO3)2 ⋅ 6H2O as metal ion sources. X-ray single-crystal diffraction analysis indicated that complex 1 contains a defect tetracubane core and possesses central symmetric structure, with two DyIII ions being in the central body position of the molecule and four CoII ions being arranged at the outer sites. Magnetic studies reveal that complex 1 behaves as single-molecule magnet (SMM) with energy barrier of 27.50 K. To investigate the individual contribution of DyIII and CoII ions to the SMM behavior, another two complexes of formulae [Dy2Zn4(L)4(NO3)2(OH)4] ⋅ 4CH3OH ( 2 ) and [Y2Co4(L)4(NO3)2(OH)4(C2H5OH)2] ⋅ 2 C2H5OH ( 3 ) were prepared. Complexes 1 and 3 are isomorphous. The coordination geometries of DyIII ions in 1 and 2 are different. The DyIII ions are eight-coordinated in 2 and nine-coordinated in 1 . Complex 2 exhibits SMM behavior with energy barrier of 69.67 K, but complex 3 does not display SMM property. These results reveal that the SMM behaviors of 1 and 2 are mainly originated from DyIII ions. It might be the higher symmetry of DyIII ions in 2 that results in the higher energy barrier.  相似文献   

2.
Cyclic sulphurylphosphazochloride (I: formula see ?Inhaltsübersicht”?) reacts with ammonia or methylamine forming the. tetramide(II) and (III), respectively, and with aniline or dimethylamine forming the diamide (IV) resp. (V). The synthesis of the diphenyl derivative (VI) is achieved starting from C6H5? PCl4. (II) gives with PCl5 the ionic compound (VII).  相似文献   

3.
The 13C NMR spectra of 15 stereoisomers of 1-mono-, 1,2-di-, and 1,2,5-trisubstituted piperidin-4-ones were investigated, and the stereochemical orientations of the substituents and the conformation relationships were established. For the series of piperidones a method for determination of the absolute confirmation of the C(2) center of the piperidone ring is proposed on the basis of the chemical shifts of the methyl groups in the 1-s--phenylethyl substituent at the nitrogen atom in the preferred rotamers with respect to the C(1)-N bond.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 11, pp. 1503–1515, November, 1986.  相似文献   

4.
The title compound, (η5‐cyclo­penta­dienyl)[(1,2,3,4,5‐η)‐4‐ferro­cenyl‐1,2,5,6‐tetrakis­(tri­methyl­silyl)­cyclo­hexa‐2,4‐dien‐1‐yl]­titanium(II), [TiFe(C5H5)2(C23H42Si4)] or [Ti{η5‐C6H2{Fe­(η5‐C5H4)(η5‐C5H5)}{Si(CH3)3}4}(η5‐C5H5)], possesses two directly linked metallocene units that subtend an angle of 52.9 (1)° (defined by the least‐squares planes of the directly connected π‐ligands) associated with the steric requirements of the bulky tri­methyl­silyl substituents. The cyclo­hexa­dienyl ligand adopts an envelope conformation; the perpendicular distance of its η5‐plane to the Ti atom is 1.512 (1) Å.  相似文献   

5.
Polybromobenzenes C6Br5X (X = Br, F, CN, NO2) react with primary amines (methylamine and cyclohexylamine) to give nucleophilic substitution products; reactions of the same substrates with secondary amines (dimethylamine, diethylamine, piperidine, and morpholine) are accompanied by hydrodebromination processes.  相似文献   

6.
By means of appearance potential measurements and metastable analysis, it is shown that the [C2H6N]+ and [CH4N]+ ions formed from methylated formamides, acetamides, ureas and thioureas rearrange upon formation to structures similar to the [M? H]+ ions from mono- and dimethylamine.  相似文献   

7.
The derivatives of the H2RuOs3(CO)13 and H4Ru4(CO)12 carbonylhydride clusters containing functionalized (including chiral) phosphines were synthesized. The solid-state structure of H2RuOs3(CO)12(Ph2P(C4H3S)) was determined by X-ray diffraction analysis. The structures of other compounds in solution were determined using IR and 1H and 31P NMR spectroscopy. A study of the temperature dependences of the 1H NMR spectra of the phosphine-substituted tetrahedral clusters along with the analysis of literature data for their analogs showed that compounds of this type exist in solution as an equilibrium mixture of isomers, which differ in arrangement of the hydride ligands at the cluster skeleton. Interconversion of the isomeric forms is due to migration of the hydride ligands over the cluster skeleton. A general model for this dynamic process was proposed. The model is consistent with both our data and earlier results of other authors. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1294–1301, July, 2007.  相似文献   

8.
Two types of manganese complexes with [Mn4] cores featuring the unusual distorted cube topology are presented, the first of which comprises new modifications of the reported complex [MnIII4(sao)4(saoH)4]·3CHCl3: [Mn4(sao)4(saoH)4]·1.32(C4H10O)·0.43(CH4O) ( 1a ) and [Mn(sao)4(saoH)4]·0.5(CH4O)·0.5(C2H3N) ( 1b ) sao = salicylaldoxime. The second, 0.55[Mn4Cl4(C12H9N2O)4(CH3OH)2(H2O)2]·0.45[Mn4Cl4(C12H9N2O)4(CH3OH)4] ( 2 ), is the first reported case of a {MnII4} core of this topology besides known {MnIII4} compounds. Differences between the {MnII4} and {MnIII4} situation are discussed, and so far overlooked differences in magnetic properties between different {MnIII4} compounds are pointed out.  相似文献   

9.
The carboxylate compounds [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})(O2CCH2SXyl)2] (2; Xyl = 3,5‐Me2C6H3) and [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})(O2CCH2SMesl)2] (3; Mes 1 = 2,4,6‐Me3C6H2) were synthesized by the reaction of [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})Cl2] (1) with 2 equivalents of xylylthioacetic acid or mesitylthioacetic acid, respectively. Compounds 2 and 3 were characterized by spectroscopic methods. The cytotoxic activity of 1–3 was tested against human tumor cell lines from four different histogenic origins—8505C (anaplastic thyroid cancer), DLD‐1 (colon cancer) and the cisplatin sensitive A253 (head and neck cancer) and A549 (lung carcinoma)—and compared with those of the reference complex [Ti(η5‐C5H5)2Cl2] (R1) and cisplatin. Surprisingly, the cytotoxic activities of the carboxylate derivatives were lower than those of their corresponding dichloride analogue (1). However, complexes 1–3 were more active than titanocene dichloride against all the studied cells with the exception of complex 2 against A253 and A549 cell lines. DNA‐interaction tests were also carried out. Solutions of all the studied complexes were treated with different concentrations of fish sperm DNA, observing modifications of the UV spectra with intrinsic binding constants of 2.99 × 105, 2.45 × 105, and 2.35 × 105 M ?1 for 1–3. Structural studies based on density functional theory calculations of 2 and 3 were also carried out. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
The title compounds, [Fe(C5H5)(C14H13O2)] and [Fe(C5H5)(C15H15O2)], respectively, contain the ferrocenyl η5(C5H4) and phenyl­ene –C6H4– rings in a nearly coplanar arrangement, with interplanar angles of 6.88 (12) and 10.5 (2)°, respectively. Molecules of the ethyl ester form dimers through η5(C5H5)C—H⋯O=C hydrogen bonds, with graph set R(20), and, together with Csp3—H⋯π(C5H5) interactions, generate a one‐dimensional column (irregular ladder). Molecules of the iso­propyl ester aggregate through η5(C5H5)C—H⋯π(C6H4) interactions.  相似文献   

11.
The reactions of glyoxal (CHO)2) with amines in cloud processes contribute to the formation of brown carbon and oligomer particles in the atmosphere. However, their molecular mechanisms remain unknown. Herein, we investigate the ammonolysis mechanisms of glyoxal with amines at the air-water nanodroplet interface. We identified three and two distinct pathways for the ammonolysis of glyoxal with dimethylamine and methylamine by using metadynamics simulations at the air-water nanodroplet interface, respectively. Notably, the stepwise pathways mediated by the water dimer for the reactions of glyoxal with dimethylamine and methylamine display the lowest free energy barriers of 3.6 and 4.9 kcal ⋅ mol−1, respectively. These results showed that the air-water nanodroplet ammonolysis reactions of glyoxal with dimethylamine and methylamine were more feasible and occurred at faster rates than the corresponding gas phase ammonolysis, the OH+(CHO)2 reaction, and the aqueous phase reaction of glyoxal, leading to the dominant removal of glyoxal. Our results provide new and important insight into the reactions between carbonyl compounds and amines, which are crucial in forming nitrogen-containing aerosol particles.  相似文献   

12.
Fourier transform ion cyclotron resonance mass spectrometry has been used to measure the reaction rates for ions derived from methylamine with dimethylamine or trimethylamine. The use of the selective ion ejection technique greatly simplifies the elucidation of the ion-molecule reaction channels. The rate constants for proton transfer from protonated metwlamine, CH3NH 3 + (m/z 32), to dimethylamine and trimethylamine are 16.1 ± 1.6 × 10?10 and 9.3 ± 0.9 × 10?10 cm3 molec?1s?1, respectively. The rate constants for charge transfer from methylamine molecular ion, CH3NH 2 + (m/z 31), to dimethylamine and trimethylamine are 9.3 ± 1.8 x 10?10 and 15.0 ± 5 × 10?10 cm3molec?1s?1, respectively.  相似文献   

13.
The dirhodium complex bis­(benzonitrile)tetra­kis[μ‐4‐(diethyl­amino)benzoato‐κ2O:O′]dirhodium(II)(RhRh) benzonitrile disolvate, [Rh2(C11H14NO2)4(C7H5N)2]·2C7H5N, lies about an inversion centre. The dirhodium complex (methanol)tetra­kis(μ‐4‐nitro­benzoato‐κ2O:O′)(pyridine)dirhodium(II)(RhRh) dichloro­methane solvate, [Rh2(C7H4NO4)4(C5H5N)(CH4O)]·CH2Cl2, lies in a general position in the unit cell, but the complexes dimerize around an inversion centre via O—H⋯O hydrogen bonding of the axial MeOH to a carboxyl­ate O atom. In the latter crystal structure, π–π stacking inter­actions between the bridging 4‐nitro­benzoate ligands and the axial pyridine ligand are observed between adjacent mol­ecules.  相似文献   

14.
Two new one‐dimensional CuII coordination polymers (CPs) containing the C2h‐symmetric terphenyl‐based dicarboxylate linker 1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylate (3,3′‐TPDC), namely catena‐poly[[bis(dimethylamine‐κN)copper(II)]‐μ‐1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylato‐κ4O,O′:O′′:O′′′] monohydrate], {[Cu(C20H12O4)(C2H7N)2]·H2O}n, (I), and catena‐poly[[aquabis(dimethylamine‐κN)copper(II)]‐μ‐1,1′:4′,1′′‐terphenyl‐3,3′‐dicarboxylato‐κ2O3:O3′] monohydrate], {[Cu(C20H12O4)(C2H7N)2(H2O)]·H2O}n, (II), were both obtained from two different methods of preparation: one reaction was performed in the presence of 1,4‐diazabicyclo[2.2.2]octane (DABCO) as a potential pillar ligand and the other was carried out in the absence of the DABCO pillar. Both reactions afforded crystals of different colours, i.e. violet plates for (I) and blue needles for (II), both of which were analysed by X‐ray crystallography. The 3,3′‐TPDC bridging ligands coordinate the CuII ions in asymmetric chelating modes in (I) and in monodenate binding modes in (II), forming one‐dimensional chains in each case. Both coordination polymers contain two coordinated dimethylamine ligands in mutually trans positions, and there is an additional aqua ligand in (II). The solvent water molecules are involved in hydrogen bonds between the one‐dimensional coordination polymer chains, forming a two‐dimensional network in (I) and a three‐dimensional network in (II).  相似文献   

15.
A chain-like zincophosphate [Zn8(HPO4)8(H2PO4)8]•[(C2H8N)8]•4H2O was obtained at room temperature from a ZnO/P2O5/dimethylamine/H2O mixture. The crystal structure was determined by single crystal X-ray diffraction. The symmetry is monoclinic a=1.26450(7)nm, b=1.08477(5)nm, c=1.46311(4)nm, β=98.793(5)°, space group Cc. The structure consists of chains of zinc-corner-sharing Zn2P2O4 four rings. The negative charge of the chains is compensated by the protonated dimethylamine. The characterization by 31P solid state nmr spectroscopy is also reported.  相似文献   

16.
The central Ge atoms in the structures of 3‐(2‐fluoro­phenyl)‐3‐(tri­phenyl­germyl)­propionic acid, [Ge(C6H5)3(C9H8FO2)], 3‐(2‐tolyl)‐3‐(tri‐4‐tolyl­germyl)­propionic acid, [Ge(C7H7)3(C10H11O2)], and 3‐(4‐tolyl)‐3‐(tri­benzyl­germyl)­propionic acid, [Ge(C7H7)3(C10H11O2)], are four‐coordinate with slightly distorted tetrahedral geometry. The Ge—Csp3 distances [1.970 (3)–1.997 (3) Å] are significantly longer than the Ge—Caromatic distances [1.940 (3)–1.959 (2) Å]. In all three structures, the mol­ecules form dimeric pairs about inversion centres through strong hydrogen‐bonding interactions between carboxyl­ic acid groups.  相似文献   

17.
[Cp°MoCl4] (Cp° = C5EtMe4) reacts with primary phosphines PH2R to give the paramagnetic phosphine complexes [Cp°MoCl4(PH2R)] [Cp° = C5EtMe4, R = But ( 1 ), 1‐Ad (1‐Ad = 1‐adamantyl; 2 ), Cy ( 3 ), Ph ( 4 ), Mes (Mes = 2, 4, 6‐Me3C6H2; 5 ), Tipp (Tipp = 2, 4, 6‐Pri3C6H2; 6 )]. 1 — 6 were characterized spectroscopically (IR, MS), and X‐ray crystal structures were determined for 1 — 4 and 6 . EPR investigations in liquid and frozen solution confirmed the presence of MoV species, and the data were used to analyze the spin‐density distribution in the first coordination sphere. Complexes 3 and 4 react with two equivalents of NEt3 with formation of [Cp°MoCl23‐P4Cy4H)] ( 7 ) and [Cp°2Mo2(μ‐Cl)2(μ‐P4Ph4)] ( 8 ), respectively, in low yield. Complexes 7 and 8 were characterized by X‐ray crystallography.  相似文献   

18.
Reaction of Mo(CO)(η2‐C2Ph2)24‐C4Ph4) and Me3NO in acetonitrile solvent affords Mo(NCMe)(η2‐C2Ph2)24‐C4Ph4) 1 . Compound 1 reacts with trimethylphosphine to produce Mo(PMe3)(η2‐C2Ph2)24‐C4Ph4) 2 , or reacts with diphenylacetylene to produce (η5‐C5Ph5)2Mo 3 and Mo(η2‐O2CPh)(η4‐C4Ph4H)(η4‐C4Ph4) 4 . The molecular structures of 1, 2 and 4 have been determined by an X‐ray diffraction study.  相似文献   

19.
[Mn(H2O)4(C4N2H4)][C6H4(COO)2] – An One‐Dimensional Coordination Polymer with Chain‐like [Mn(H2O)4(C4N2H4)]n2n+ Polycations Orthorhombic single crystals of [Mn(H2O)4(C4N2H4)][C6H4(COO)2] have been prepared in aqueous solution at room temperature. Space group Imm2 (no. 44), a = 1039.00(6) pm, b = 954.46(13) pm, c = 737.86(5) pm, V = 0.73172(12) nm3, Z = 2. Mn2+ is coordinated in a octahedral manner by four water molecules and two nitrogen atoms stemming from the pyrazine molecules (Mn–O 215.02(11) pm; Mn–N 228.7(4), 230.7(4) pm). Mn2+ and pyrazine molecules form chain‐like polycations with [Mn(H2O)4(C4N2H4)]n2n+ composition. The positive charge of the polycationic chains is compensated for by phthalate anions, which are accomodated between the chains. The phthalate anions are linked by hydrogen bonds to the polycationic chains. Thermogravimetric analysis in air revealed that the loss of water of crystallisation and pyrazine occurs in two steps between 130 and 245 °C. The resulting sample was stable up to 360 °C. Further decomposition yielded Mn2O3.  相似文献   

20.
The geometric features of 1‐(4‐nitrophenyl)‐1H‐tetrazol‐5‐amine, C7H6N6O2, correspond to the presence of the essential interaction of the 5‐amino group lone pair with the π system of the tetrazole ring. Intermolecular N—H...N and N—H...O hydrogen bonds result in the formation of infinite chains running along the [110] direction and involve centrosymmetric ring structures with motifs R22(8) and R22(20). Molecules of {(E)‐[1‐(4‐ethoxyphenyl)‐1H‐tetrazol‐5‐yl]iminomethyl}dimethylamine, C12H16N6O, are essentially flattened, which facilitates the formation of a conjugated system spanning the whole molecule. Conjugation in the azomethine N=C—N fragment results in practically the same length for the formal double and single bonds.  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