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.
Two new charge-transfer salts, [CpFeCpCH2N(CH3)3]4[PMo12O40] · CH3CN (1) and [CpFeCpCH2N(CH3)3]4[GeMo12O40] (2), were synthesized by the traditional solution synthetic method and their structures were determined by single-crystal X-ray analysis. Salt 1 belongs to the triclinic space group P1, and salt 2 belongs to the triclinic space group . There exist the complex interactions of the cationic ferrocenyl donor and Keggin polyanion in the solid state. The solid state UV-Vis diffuse reflectance spectra indicate the presence of a charge-transfer band climbing from 450 nm to well beyond 900 nm for 1, a charge-transfer band from 460 to 850 nm with λmax = 630 nm for 2.The EPR spectra of salts 1 and 2 at 77 K show a signal at g = 2.0048 and 1.9501, respectively, ascribed to the delocalization of one electron in reduced Keggin ion in salt 1 and the MoVI in [GeMo12O40]4− is partly reduced to MoV owing to the charge-transfer transitions taking place between the ferrocenyl donors and the POM acceptors. The two compounds were also characterized by IR spectroscopy and cyclic voltammetry.  相似文献   

2.
The crystalline compounds [AlMen{Si(SiMe3)3}3−n(thf)] [n = 2 (1) or 1 (2)] were prepared from the lithium sisyl [Li{Si(SiMe3)3}(thf)3] (A) and the appropriate methylaluminium chloride [AlCl3−nMen] in thf. The X-ray structure of 1 is reported. Unlike A or a magnesium sisyl [Mg{Si(SiMe3)3}2(thf)2] (B), neither 1 nor 2 underwent an insertion reaction with an α-H-free nitrile.  相似文献   

3.
New stable heteroleptic germanium(II) and tin(II) compounds [(SiMe3)2N-E14-OCH2CH2NMe2]n (E14 = Ge, n = 1 (1), Sn, n = 2 (2)) have been synthesized and their crystal structures have been determined by X-ray diffraction analysis. While compound 1 is monomer stabilized by intramolecular Ge ← N coordination, compound 2 is associated to dimer via intermolecular dative Sn ← O interactions.  相似文献   

4.
Quantum chemical calculations using DFT at the B3LYP level have been carried out for the reaction of ethylene with the group-7 compounds ReO2(CH3)(CH2) (Re1), TcO2(CH3)(CH2) (Tc1) and MnO2(CH3)(CH2) (Mn1). The calculations suggest rather complex scenarios with numerous pathways, where the initial compounds Re1-Mn1 may either engage in cycloaddition reactions or numerous addition reactions with concomitant hydrogen migration. There are also energetically low-lying rearrangements of the starting compounds to isomers which may react with ethylene yielding further products. The [2 + 2]Re,C cycloaddition reaction of the starting molecule Re1 is kinetically and thermodynamically favored over the [3 + 2]C,O and [3 + 2]O,O cycloadditions. However, the reaction which leads to the most stable product takes place with initial rearrangement to the dioxohydridometallacyclopropane isomer Re1a that adds ethylene with concomitant hydrogen migration yielding Re1a-1. The latter reaction has a slightly higher barrier than the [2 + 2]Re,C cycloaddition reaction. The direct [3 + 2]C,O cycloaddition becomes more favorable than the [2 + 2]M,C reaction for the starting compounds Tc1 and Mn1 of the lighter metals technetium and manganese but the calculations predict that other reactions are kinetically and thermodynamically more favorable than the cycloadditions. The reactions with the lowest activation barriers lead after rearrangement to the ethyl substituted dioxometallacyclopropanes Tc1a-1 and Mn1a-1. The manganese compound exhibits an even more complex reaction scenario than the technetium compounds. The thermodynamically most stable final product of ethylene addition to Mn1 is the ethoxy substituted metallacyclopropane Mn1a-2 which has, however, a high activation barrier.  相似文献   

5.
The reactions of the sterically demanding group-13 alkyls ER3 (E = Al, Ga, In; R = CH2t-Bu, CH2SiMe3) with the platinum-complex [(dcpe)Pt(H)(CH2t-Bu)] were re-investigated. The bimetallic compounds [(dcpe)Pt(ER2)(R)] (3: E = Ga, R = CH2SiMe3; 5: E = In, R = CH2t-Bu; dcpe = bis(dicyclohexylphosphino)ethane) with direct σ(Pt-E) bonds were obtained by oxidative addition of an E-C bond to the coordinatively unsaturated fragment [(dcpe)Pt] produced in situ by thermolysis of the starting complex [(dcpe)Pt(CH2t-Bu)(H)]. The single crystal structure determination reveals a Pt-Ga bond length of 2.376(2) Å and a Pt-In bond length of 2.608(1) Å. All new compounds were characterised by elemental analysis, 31P and 195Pt NMR spectroscopy. Interestingly, the Pt-Ga compound 3 slowly transforms into the platinum alkyl/hydride isomer {(dcpe)Pt(μ2-H)[CH2Si(CH3)2 CH2Ga(CH2SiMe3)2]} (4) during crystallization from solution at room temperature. The X-ray single crystal structure analysis revealed both complexes 3 and 4 coexisting in the unit cell in a 1:1 relation. The inaccessibility of analytically pure samples of the Pt-Al complex {(dcpe)Pt[Al(CH2t-Bu)2](CH2t-Bu)} (6), postulated as intermediate of the reaction of [(dcpe)Pt(H)(CH2t-Bu)] with Al(CH2t-Bu) on the basis of 31P and 195Pt NMR data, is attributed to an enhanced tendency to isomerisation into the alkyl/hydride Pt/Al congener of 4. A brief DFT analysis of the bonding situation of the model complex [(dhpe)Pt(GaMe2)(Me)] (1M) revealed, that the contribution of π(Pt-Ga) back-bonding is negligible.  相似文献   

6.
The new mixed Sb2O-donor ligands O{(CH2)2SbR2}2 (R = Ph, 1; R = Me, 2) with flexible backbones have been prepared in good yields as air-sensitive oils from reaction of NaSbR2 with 0.5 mol equivalents of O(CH2CH2Br)2 in thf solution. The As2O-donor analogues, O{(CH2)2AsR2}2 (R = Ph, 3; R = Me, 4) were obtained similarly from LiAsPh2 or NaAsMe2, respectively and O(CH2CH2Br)2, although ligand 4 appears to be considerably less stable with respect to C-O bond fission under some conditions than the other ligands. Using O(CH2CH2Cl)2 leads only to partial substitution by the SbPh2 or AsPh2 nucleophile. These ligands behave as bidentate chelating Sb2- or As2-donors in the distorted tetrahedral [M(L-L)2]BF4 (M = Cu or Ag; L-L = 1-4) on the basis of solution 1H and 63Cu NMR spectroscopic studies, mass spectrometry and microanalyses. Crystal structures of three representative examples with Cu(I) and Ag(I) confirm the distorted tetrahedral Sb4 or As4 coordination at the metal and allow comparisons of geometric parameters. The crystallographic identification of an unexpected Cu(I)-Cu(I) complex, [Cu2{Me2As(CH2)2OH}3](BF4)2, obtained as a by-product via C-O bond fission within ligand 4 is also reported. The distorted octahedral [RhCl2(L-L)2]Cl and the distorted square planar cis-[PtCl2(L-L)] (L-L = 1 or 2) are also described. The ether O atoms are not involved in coordination to the metal ion in any of the late transition metal complexes isolated.  相似文献   

7.
The reaction between ClCH2-R-CH2Cl, R = p-C6H4, and [Ph3Sn]Li+ yields Ph3Sn-CH2-R-CH2-SnPh3 (1) in high yield. The related known compound R = CH2CH2 (1a) is synthesized by the reaction of the di-Grignard reagent BrMg(CH2)4MgBr with two equivalents of Ph3SnCl. Cleavage of a single Sn-Ph group at each tin centre of both compounds using HCl/Et2O yields the corresponding bis-chlorostannanes Ph2ClSn-CH2-R-CH2-SnClPh2, R = (CH2)4 (2) and R = C6H4 (3), respectively. Compounds 1, 2 and 3 are crystalline solid materials and their single crystal X-ray structures are reported. In the solid state both 2 and 3 form self-assembled ladder structures involving alternating intermolecular Cl-Sn?Cl and Cl?Sn-Cl bonded chains at both ends of the distannanes with 5-coordinate tin atoms. Recrystallization of 3 from CH2Cl2 in the presence of DMF yields the bis-DMF adduct (4) in which no self-assembled structures were noted. Evaluation of the chlorostannanes 2 and 3 against a suite of bacteria, Staphylococcus aureus, Escherichia coli and Photobacterium phosphoreum is reported and compared to the related mono-chlorostannanes Ph2(CH3)SnCl and Ph2(PhCH2)SnCl.  相似文献   

8.
A versatile neutral metalloligand [Cu(PySal)2] (1) (PySal = 3-pyridylmethylsalicylidene-imino) was exploited as a building unit to construct five complexes {Cu[Cu(PySal)2]2}(ClO4)2 (2), {Cd[Cu(PySal)2]2(H2O)2]} (NO3)2 · 2H2O · 4CH3OH (3), {Zn[μ2-Cu(PySal)2]Cl2}n · nCH3OH (4), {Hg[μ2-Cu(PySal)2]I2}n (5) and {Cd[μ2-Cu(PySal)2]Cl2}n · nCH2Cl2 (6). [Cu(PySal)2] acts as a chelating ligand in discrete complexes 2 and 3 with unbound anions, but as a bis-monodentate bridging ligand in polymers 4, 5 and 6 when halogen anions coordinated cooperatively to metal cations. The coordination geometry of Cu2+ is well-defined square planar in bridging [Cu(PySal)2], analogous to that in free metalloligand (1), but it is distorted square planar in chelating [Cu(PySal)2].  相似文献   

9.
A series of new compounds containing rare earth cations (Eu to Yb) and paramagnetic cluster anion [Re6Te8(CN)6]3− was prepared and investigated. The X-ray structural analyses have revealed that the compounds [{Ln(H2O)4}{Re6Te8(CN)6}] · 2.5H2O; Ln = Eu (1), Tb (3), Dy (4), Ho (5), Er (6), Tm (7), [{Gd(H2O)3}{Re6Te8(CN)6}] · 2.5H2O (2) and [{Yb(H2O)4}{Re6Te8(CN)6}] (8) are three-dimensional polymers based on Re–CN–Ln interactions. Measurements of magnetic susceptibility for 2 and 5 showed that effective magnetic moment (at 300 K) was 8.13 μB for compound 2 and 10.79 μB for compound 5 with weak antiferromagnetic ordering appeared at low temperatures.  相似文献   

10.
Ph2SiCl2 and PhMeSiCl2 react with Li2E (E = S, Se, Te) under formation of trimeric diorganosilicon chalcogenides (PhRSiE)3 (R = Ph: 1a-3a, R = Me: cis/trans-4a (E = S), cis/trans-5a (E = Se)). In case of E = S, Se dimeric four-membered ring compounds (PhRSiE)2 (R = Ph: 1b-2b, R = Me: cis/trans-4b (E = S), cis/trans-5b (E = Se)) have been observed as by-products. 1a-5b have been characterized by multinuclear NMR spectroscopy (1H, 13C, 29Si, 77Se, 125Te). Four- and six-membered ring compounds differ significantly in 29Si and 77Se chemical shifts as well as in the value of 1JSiSe.The molecular structures of 2a, 3a and trans-5a reported in this paper are the first examples of compounds with unfused six-membered rings Si3E3 (E = Se, Te). The Si3E3 rings adopt twisted boat conformations. The crystal structure of 3a reveals an intermolecular Te-Te contact of 3.858 Å which yields a dimerization in the solid state.  相似文献   

11.
The reaction of HgCl2 and Te(R)CH2SiMe3 [R = CH2SiMe3 (1), Ph (2)] in ethanol yielded a mononuclear complex [HgCl2{Te(R)CH2SiMe3}2] (R = Ph, 3a; R = CH2SiMe3, 3b). The recrystallization of 3a or 3b from CH2Cl2 produced a dinuclear complex [Hg2Cl2(μ-Cl)2{Te(R)CH2SiMe3}2] (R = Ph, 4a; R = CH2SiMe3, 4b). When 3a was dissolved in CH2Cl2, the solvent quickly removed, and the solid recrystallized from EtOH, a stable ionic [HgCl{Te(Ph)CH2SiMe3}3]Cl·2EtOH (5a·2EtOH) was obtained. Crystals of [HgCl2{Te(CH2SiMe)2}]·2HgCl2·CH2Cl2 (6b·2HgCl2·CH2Cl2) were obtained from the CH2Cl2 solution of 3b upon prolonged standing. The complex formation was monitored by 125Te-, and 199Hg NMR spectroscopy, and the crystal structures of the complexes were determined by single crystal X-ray crystallography.  相似文献   

12.
A series of organotin compounds bearing two intramolecular N → Sn coordination bonds RSn(OCH2CH2NMe2)2Cl (R = Me (4), n-Bu (5), Mes (6)) were synthesized in good yields. These compounds as well as 2 (R = Ph) react with PhSnCl3 to give redistribution products RPhSnCl2 and (Me2NCH2CH2O)2SnCl2 (3). The direction of redistribution reactions is reverse to Kocheshkov reaction. DFT calculations have shown that the driving force of the reactions is formation of intramolecular N → Sn coordination bonds in (RO)2SnCl2 (3), the Lewis acid stronger than RSn(OR)2Cl (2, 4-6). The mechanism of the redistribution reaction between 2 and PhSnCl3 consists of two steps: (1) initial exchange of OCH2CH2NMe2 and Cl to give PhSn(OCH2CH2NMe2)Cl2 (7) followed by (2). Ph and OCH2CH2NMe2 exchange.  相似文献   

13.
Complexes M(CCCSiMe3)(CO)2Tp′ (Tp′ = Tp [HB(pz)3], M = Mo 2, W 4; Tp′ = Tp [HB(dmpz)3], M = Mo 3) are obtained from M(CCCSiMe3)(O2CCF3)(CO)2(tmeda) (1) and K[Tp′].Reactions of 2 or 4 with AuCl(PPh3)/K2CO3 in MeOH afforded M{CCCAu(PPh3)}(CO)2Tp′ (M = Mo 5, W 6) containing C3 chains linking the Group 6 metal and gold centres.In turn, the gold complexes react with Co33-CBr)(μ-dppm)(CO)7 to give the C4-bridged {Tp(OC)2M}CCCC{Co3(μ-dppm)(CO)7} (M = Mo 7, W 8), while Mo(CBr)(CO)2Tp and Co33-C(CC)2Au(PPh3)}(μ-dppm)(CO)7 give {Tp(OC)2Mo}C(CC)2C{Co3(μ-dppm)(CO)7} (9) via a phosphine-gold(I) halide elimination reaction. The C3 complexes Tp′(OC)2MCCCRu(dppe)Cp (Tp′ = Tp, M = Mo 10, W 11; Tp′ = Tp, M = Mo 12) were obtained from 2-4 and RuCl(dppe)Cp via KF-induced metalla-desilylation reactions. Reactions between Mo(CBr)(CO)2Tp and Ru{(CC)nAu(PPh3)}(dppe)Cp (n = 2, 3) afforded {Tp(OC)2Mo}C(CC)n{Ru(dppe)Cp} (n = 2 13, 3 14), containing C5 and C7 chains, respectively. Single-crystal X-ray structure determinations of 1, 2, 7, 8, 9 and 12 are reported.  相似文献   

14.
The reaction of the anion [(tBuP)3As] (1) with Me2SiCl2 results in nucleophilic substitution of the Cl anions, giving the di- and mono-substituted products [Me2Si{As(PtBu)3}2] (3a) and [Me2Si(Cl){As(PtBu)3}] (3b). Analogous reactions of the pre-isolated [(CyP)4As] anion (2) (Cy = cyclohexyl) with Me2SiCl2 produced mixtures of products, from which no pure materials could be isolated. However, reaction of 2 [generated in situ from CyPHLi and As(NMe2)3] gives the heterocycle [(CyP)3SiMe2] (4). The X-ray structures of 3a and 4 are reported.  相似文献   

15.
Bis(dimethylphosphino)naphthalene, 1,8-(PMe2)2C10H6 (dmpn), reacts readily with Ru3(CO)12 or Ru3(μ-dppm)(CO)10 with replacement of one of the PMe2 groups by H to give Ru3(CO)12 − n{PMe2(nap)}n (n = 1 2, 2 3) or Ru3(μ-dppm)(CO)9{PMe2(nap)} 4; the complex Ru3(CO)10(dmpn) 1 is obtained only in small amount. Thermolysis of 2 or 4 gives Ru3(μ-H)23-PMe2(C10H5)}(μ-dppm)n (CO)8-2n (n = 0 5, 1 6, respectively) containing μ3-naphthalyne groups.  相似文献   

16.
The energies and structures of possible intermediates in the dinitrogen extrusion from diazidophenylborane 4a to give phenylborylene 11a were determined using density functional (B3LYP), multiconfigurational (CASSCF and MRMP2), and coupled cluster (CCSD(T)) computations in conjunction with basis sets of up to cc-pVTZ quality. Formation of 11a and nitrogen from 4a is an exothermic process (−21 kcal mol−1). The triplet electronic ground state of azidophenylborylnitrene 5a (PhBN4) is only 26 kcal mol−1 higher in energy than 4a and the phenyl shift in 5a to yield N-azidophenyliminoborane 7a is highly exothermic.  相似文献   

17.
The crystal structures of (2-aza-2-benzyl-5,10,15,20-tetraphenyl-21-carbaporphyrinato-N,N′,N″) nickel(II) methylene chloride solvate [Ni(2-NCH2C6H5NCTPP); 4], (2-aza-2-benzyl-5,10,15,20-tetraphenyl-21-carbaporphyrinato-N,N′,N″) palladium(II) [Pd(2-NCH2C6H5NCTPP); 5] and bromo(2-aza-2-benzyl-5,10,15,20-tetraphenyl-21-carbaporphyrinato-N,N′,N″) manganese(III) toluene solvate [Mn(2-NCH2C6H5NCTPP)Br·C6H5CH3; 3·C6H5CH3] have been established. The coordination sphere around the Ni2+ ion in 4 (or Pd2+ ion in 5) is distorted square planar (DSP), whereas for Mn3+ in 3·C6H5CH3, it is a square-based pyramid with the Br atom lying in the axial site. The g value of 11.34, measured from parallel polarization of the X-band EPR spectra at 4 K, is consistent with a high spin mononuclear manganese(III) centre (S = 2) in 3. The magnitude of the axial (D) zero-field splitting (ZFS) for the mononuclear Mn(III) centre in 3 was determined approximately to be 1.4 cm−1 by paramagnetic susceptibility measurements and conventional EPR spectroscopy.  相似文献   

18.
A series of heterobinuclear ferrocene-ruthenium complexes Fc(CHCH)nRuCl(CO)(PMe3)3 (n = 1, 3; n = 2, 12), Fc(CHCH)RuCl(CO)(Py)(PPh3)2 (4), and trimetallic Fc(CHCH)RuCl(CO)(PPh3)2(Py-E-(CHCH)Fc) (6) have been prepared. The length of the molecular rods is extended by successive insertion of CHCH spacers in the bridging ligands or the ancillary ligands. The respective products have been fully characterized and the structures of 3 and 12 have been established by X-ray crystallography. Electrochemical studies have revealed that ethenyl heterobimetallic complexes display two successive one-electron processes, and that intermetallic electronic communication between the two endgroups is attenuated with the increase of the length of the conjugated bridge. The electrochemical behavior of the trimetallic complex reveals strong electronic communication between ruthenium and ferrocene transmitted through the ethenyl bridge, however, it also reveals a very weak interaction between ruthenium and ferrocene transmitted through the (E)-CHCH-Py bridge.  相似文献   

19.
We describe reactions of [99mTc(H2O)3(CO)3)]+ (1) with Diels-Alder products of cyclopentadiene such as “Thiele’s acid” (HCp-COOH)2 (2) and derivatives thereof in which the corresponding [(Cp-COOH)99mTc(CO)3)] (3) complex did form in water. We propose a metal mediated Diels-Alder reaction mechanism. To show that this reaction was not limited to carboxylate groups, we synthesized conjugates of 2 (HCp-CONHR)2 (4a-c) (4a, R = benzyl amine; 4b, R = Nα-Boc-l-2,3-diaminopropionic acid and 4c, R = glycine). The corresponding 99mTc complexes [(4a)99mTc(CO)3)] 6a, [(4b)99mTc(CO)3)] 6b and [(4c)99mTc(CO)3)] 6c have been prepared along the same route as for Thiele’s acid in aqueous media demonstrating the general applicability of this synthetic strategy. The authenticity of the 99mTc complexes on the no carrier added level have been confirmed by chromatographic comparison with the structurally characterized manganese or rhenium complexes.Studies of the reaction of 1 with Thiele’s acid bound to a solid phase resin demonstrated the formation of [(Cp-COOH)99mTc(CO)3)] 3 in a heterogeneous reaction. This is the first evidence for the formation of no carrier added 99mTc radiopharmaceuticals containing cyclopentadienyl ligands via solid phase syntheses. Macroscopically, the manganese analogue 5a and the rhenium complexes 5b-c have been prepared and characterized by IR, NMR, ESI-MS and X-ray crystallography for 5a (monoclinic, P21/c, a = 9.8696(2) Å, b = 25.8533(4) Å, c = 11.8414(2) Å, β = 98.7322(17)°) in order to unambiguously assign the authenticity of the corresponding 99mTc complexes.  相似文献   

20.
Three complexes of composition [CrL(X)3], where L = 4′-(2-pyridyl)-2,2′:6′,2″-terpyridine and X = Cl, N3, NCS are synthesized. They are characterized by IR, UV–Vis, fluorescence, EPR spectroscopic, and X-ray crystallographic studies. Structural studies reveal that the Cr(III) ion is coordinated by three N atoms of L in a meridional fashion. The three anions occupy the other three coordination sites completing the mer-N3Cl3 (1) and mer-N3N3 (2 and 3), distorted octahedral geometry. The Cr–N2 has a shorter length than the Cr–N1 and Cr–N3 distances and the order Cr–N(NCS) < Cr–N(N3) < Cr–Cl is observed. They exhibit some of the d–d transitions in the visible and intra-ligand transitions in the UV regions. The lowest energy d–d transition follows the trend [CrLCl3] < [CrL(N3)3] < [CrL(NCS)3] consistent with the spectrochemical series. In DMF, they exhibit fluorescence having π → π character. All the complexes show a rhombic splitting as well as zero-field splitting (zfs) in X-band EPR spectra at 77 K.  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