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 46 毫秒
1.
A novel dinuclear copper(II) complex with the amino acid l-arginine (l-arg), with mono and bidentate HPO42− oxoanions and an OH anion. [Cu2(l-arg)2(μ-HPO4-O)(μ-HPO4-O,O′)(μ-OH)] · (H3O)+ · 6H2O (1) was prepared and its structure was determined by X-ray diffraction methods. The two independent copper ions are in a distorted square pyramidal coordination, each bonded to one l-arginine molecule. These two Cu(l-arg) units are bridged by two monoatomic equatorial–apical oxygen ligands belonging to a monodentate hydrogenphosphate group, and to the hydroxyl group. The copper ions in the dinuclear unit at d = 3.1948(8) Å are also connected by two equatorial oxygen belonging to a bidentate hydrogenphosphate. This dinuclear character and bridging scheme, not common for metal–amino acid compounds, is a consequence of the properties of the phosphate anions. The magnetic susceptibility at temperatures between 2 and 300 K and the isothermal magnetization curves at T = 2.29(1) K with applied fields up to 9 T were measured. The magnetic data indicate an antiferromagnetic intradinuclear exchange coupling J/kB = −3.7(1) K and using a molecular field approximation we estimated a weaker ferromagnetic interaction J′/kB ∼ 0.3 K between neighbour dinuclear units.  相似文献   

2.
3.
4.
The crystal structure of copper sulfate templated by 2-methylpiperazine, (C5H14N2)[Cu(SO4)2(H2O)4] · H2O, was investigated using single crystal X-ray diffraction data. At room temperature, it crystallises in the monoclinic P21/n space group with the following unit-cell parameters: a = 6.9153(1), b = 23.1295(3), c = 10.4472(1) Å, β = 104.227(1)°, V = 1619.75(4) Å3 and Z = 4. The CuII cation adopts a slightly distorted octahedral geometry, arising from four water molecules and two sulfate tetrahedra leading to the formation of [Cu(SO4)2(H2O)4] units. The structure consists of isolated [Cu(SO4)2(H2O)4]2− anions, 2-methylpiperazinediium cations (C5H14N2)2+ and water molecules connected by a three-dimensional hydrogen-bond network. The thermal decomposition of the precursor, studied by thermogravimetry and temperature-dependent X-ray powder diffraction, proceeds through four stages giving rise to the copper oxide.  相似文献   

5.
The syntheses, physical characterization and crystal structures of two new molecular copper(II) complexes of composition [Cu(C5H5N)2(C7F5O2)2] (1) and [Cu(C5H5N)2(C7F5O2)2(H2O)] (2) (C5H5N = py = pyridine and C7F5O2 = pfb = pentafluorobenzoate) are reported. Single-crystal X-ray structure determinations revealed that in 1, the Cu2+ ion, which lies on a crystallographic inversion centre, is coordinated to two py molecules and two oxygen atoms from two monodentate pfb anions, resulting in a trans-CuN2O2 square planar geometry. In 2, the Cu2+ ion is also coordinated to two py and two pfb species in addition to a water molecule in the apical site of a distorted CuN2O3 square pyramid. In the crystal packing, both 1 and 2 show segregated aromatic π-π stacking interactions in which (py + py) and (pfb + pfb) ring-pairings are seen, but no (py + pfb) pairings occur. Crystal data: 1: C24H10CuF10N2O4, Mr = 643.88, space group , a = 8.0777 (3) Å, b = 8.0937 (3) Å, c = 10.5045 (5) Å, α = 90.916 (3)°, β = 93.189 (2)°, γ = 118.245 (3)°, V = 603.36 (4) Å3, Z = 1. 2: C24H12CuF10N2O5, Mr = 661.90, space group , a = 7.5913 (5) Å, b = 15.6517 (6) Å, c = 21.1820 (14) Å, α = 95.697 (4)°, β = 94.506 (2)°, γ = 91.492 (4)°, V = 2495.2 (3) Å3, Z = 4.  相似文献   

6.
The three-dimensional coordination polymers [Ni4(μ-H2O)2(nic)8 · 2H2O] (nic = nicotinate, 3-pyridylcarboxylate) (1) and [Ni2(H2O)2(nic)4(4,4′-bpy)] (2) were prepared by the hydrothermal reaction of nickel(II) chloride, nicotinic acid, sodium hydroxide and an organoimine (several choices for 1, 4,4′-bipyridine for 2). The non-centrosymmetric crystal structure of 1 is constructed from binuclear [Ni2(μ-H2O)(μ3-nic)2]2+ subunits joined into 3-D via μ2- and μ3-nicotinate ligands, forming “bird”-shaped cavities that contain water molecule dimers. The crystal structure of 1 is compared and contrasted to two previously reported nickel(II) nicotinate phases. In contrast, the crystal structure of 2 is assembled from neutral [Ni(H2O)(μ2-nic)2] layers, connected into 3-D via tethering 4,4′-bpy moieties. 1 exhibits weak antiferromagnetic coupling across its binuclear subunits (J = −1.61(2) cm−1 for g = 2.233(2)), although anisotropy due to single-ion zero-field-splitting (D) cannot be excluded. The 3-D structures of 1 and 2 remain stable above 300 °C and 200 °C, respectively.  相似文献   

7.
The magnetic properties of α-Cu(dca)2(pyz) were examined by magnetic susceptibility, magnetization, inelastic neutron scattering (INS), muon-spin relaxation (μSR) measurements and by first-principles density functional theoretical (DFT) calculations and quantum Monte Carlo (QMC) simulations. The χ versus T curve shows a broad maximum at 3.5 K, and the data between 2 and 300 K is well described by an S = 1/2 Heisenberg uniform chain model with g = 2.152(1) and J/k= −5.4(1) K. μSR measurements, conducted down to 0.02 K and as a function of longitudinal magnetic field, show no oscillations in the muon asymmetry function A(t). This evidence, together with the lack of spin wave formation as gleaned from INS data, suggests that no long-range magnetic order takes place in α-Cu(dca)2(pyz) down to the lowest measured temperatures. Electronic structure calculations further show that the spin exchange is significant only along the Cu–pyz–Cu chains, such that α-Cu(dca)2(pyz) can be described by a Heisenberg antiferromagnetic chain model. Further support for this comes from the M versus B curve, which is strongly concave owing to the reduced spin dimensionality. α-Cu(dca)2(pyz) is a molecular analogue of KCuF3 owing to dx2-y2dx2-y2 orbital ordering where nearest-neighbor magnetic orbital planes of the Cu2+ sites are orthogonal in the planes perpendicular to the Cu–pyz–Cu chains.  相似文献   

8.
In the treatment of cyclometallated dimer [Pd(dmba)(μ-Cl)]2 (dmba = N,N-dimethylbenzylamine) with AgNO3 and acetonitrile the result was the monomeric cationic precursor [Pd(dmba)(NCMe)2](NO3) (NCMe = acetonitrile) (1). Compound 1 reacted with m-nitroaniline (m-NAN) and pirazine (pz), originating [Pd(dmba)(ONO2)(m-NAN)] (2) and [{Pd(dmba)(ONO2)}2(μ-pz)] · H2O (3), respectively. These compounds were characterized by elemental analysis, IR and NMR spectroscopy. The IR spectra of (23) display typical bands of monodentade O-bonded nitrate groups, whereas the NMR data of 3 are consistent with the presence of bridging pyrazine ligands. The structure of compound 3 was determined by X-ray diffraction analysis. This packing consists of a supramolecular chain formed by hydrogen bonding between the water molecule and nitrato ligands of two consecutive [Pd2(dmba)2(ONO2)2(μ-pz)] units.  相似文献   

9.
The new polynuclear heterometal alkoxide clusters Ln2Na8(OCH2CF3)14(THF)6 (Ln = Sm 1, Y 2, Yb 3) have been synthesized by the reaction of anhydrous LnCl3 with 7 equiv. of NaOCH2CF3 in 68–75% yields. Crystal structural analysis revealed clusters 13 are isomorphous composed of two cubanes and a double open cubane, with one face of an Ln1Na2O4 open cubane capped by an additional Ln1O2 layer. Clusters 13 show extremely high activity for the polymerization of ε-caprolactone (ε-CL) and trimethylene carbonate (TMC). The reactivity is much higher than those found for the monometallic alkoxides lanthanide complexes previously reported. The dependence of catalytic activity on lanthanide metals is observed: Yb ≈ Y < Sm for ε-CL and Yb < Y < Sm for TMC. The polymers obtained with these clusters all show a unimodal molecular weight distribution with moderate molecular weight distributions (Mw/Mn = 1.4–1.7), indicating that clusters 13 can really be used as single-component catalysts. The bimetallic cooperation and the coordination–insertion mechanism were proposed.  相似文献   

10.
Mononuclear compounds M(CO)23-C3H5)(en)(X) (X = Br, M = Mo(1), W(2); X = N3, M = Mo(3), W(4); X = CN, M = Mo(5), W(6)) and cyanide-bridged bimetallic compounds [(en)(η3-C3H5)(CO)2M(μ-CN)M(CO)23-C3H5)(en)]Br (M = Mo (7), W(8)) were prepared and characterized. These compounds are fluxional and display broad unresolved proton NMR signals at room temperature. Compounds 1-6 were characterized by NMR spectroscopy at −60 °C, which revealed isomers in solution. The major isomers of 1-4 adopt an asymmetric endo-conformation, while those of 5 and 6 were both found to possess a symmetric endo-conformation. The single crystal X-ray structures of 1-6 are consistent with the structures of the major isomer in solution at low temperature. In contrast to mononuclear terminal cyanide compounds 5 and 6, cyanide-bridged compounds 7 and 8 were found to adopt the asymmetric endo-conformation in the solid state.  相似文献   

11.
Reactions between 1,1′-(Me3SiCC)2Rc′ [Rc′ = ruthenocen-1,1′-diyl, Ru(η-C5H4-)2] and RuCl(PP)Cp′ in the presence of KF gave 1,1′-{Cp(PP)RuCC}2Rc′ [Cp′ = Cp, PP = PPh31, P(m-tol)32, dppe 3, dppf 4; Cp′ = Cp, PP = dppe 5]. Compounds 1 and 2 react with tcne to give two diastereomers a/b of the allylic (vinylcarbene) complexes 6 and 7, while methylation of 5 gave the bis-vinylidene [1,1′-{Cp(dppe)RuCCMe}2Rc′](BPh4)2 (8). The X-ray structures of 4, 6b and 8 have been determined. Cyclic voltammograms indicate that there is some electronic communication between the ruthenium end-groups through the Rc′ centre.  相似文献   

12.
The reaction of [Cu3(dppm)33-OH)](ClO4)2 (1) with heterocumulenes (XCS; X = NPh, NMe and S) has been studied. The μ3-OH ligand inserts into PhNCS and MeNCS only in the presence of methanol. Insertion products are formed in accord with earlier observations made with copper(I)-aryloxides. On heating, the insertion products convert to a S bridged cluster [Cu4(dppm)44-S)](ClO4)2 (8), having a tetrameric core. However, in the reaction with CS2, 1 is converted to 8 even at room temperature in the presence of methanol. On the other hand, the dimeric complex [Cu2(dppm)2(CH3CN)4](ClO4)2, reacts with CS2 to give (diphenylphosphinomethyl)-diphenylphosphine sulfide, Ph2P-CH2-P(S)Ph2 (dppmS), which forms the complex [Cu(dppmS)2]ClO4 (9). A single crystal X-ray crystallographic study of 9, the first copper(I) complex of dppmS has been taken up to confirm the mono-oxidation of the dppm ligand and the nuclearity of the complex. Reactions of complex 1 with heterocumulenes and with elemental sulfur, are compared.  相似文献   

13.
Alkylation of PdCl2(dotpm) (dotpm = bis(di-ortho-tolylphosphino)methane) with n-butyllithium produces the binuclear Pd(0) complex Pd2(μ-dotpm)2 and the elimination byproducts 1-butene, cis-2-butene, trans-2-butene, butane, and octane. The dibutyl complex, Pd(dotpm)(n-Bu)2, is presumed to be the reaction intermediate. The crystal structure of Pd2(μ-dotpm)2 reveals that the methylene groups of the bridging dotpm ligands are located on opposite sides of the Pd2P4 unit, forming an 8-membered ring that is in an elongated chair conformation. The four phosphorus atoms are not coplanar, and the P1-P2-P3-P4 ring has a torsion angle of 13.8°, which minimizes the spatial interactions among the o-tolyl rings. The Pd-Pd bond distance is 2.8560(6) Å, which indicates that there is a weak “closed-shell” bonding interaction between the d10-d10 metal centers. Each palladium atom has a nearly linear geometry, and the eight methyl groups of the dotpm ligands shield the open coordination sites on the metal centers. Four methyl groups shield the metal atoms above and below the Pd2P4 ring cavity, and four methyl groups block the open metal sites outside of the Pd2P4 ring. The Pd2(μ-dotpm)2 complex readily undergoes oxidative addition of dichloromethane to form the rigid A-frame complex Pd2Cl2(μ-CH2)(μ-dotpm)2.  相似文献   

14.
Compounds M(CO)23-C3H5)(L-L)(NCBH3) (L-L = dppe, M = Mo(1), W(2); L-L = bipy, M = Mo(3), W(4); L-L = en, M = Mo(5), W(6)) were prepared and characterized. The single crystal X-ray analyses of 2-6 revealed that the cyanotrihydroborate anion bonds to the metal through a nitrogen atom, the open face of the allyl group being pointed toward the two carbonyls (endo-isomer). In compounds 2, 5, and 6, the two donor atoms of the bidentate ligand occupy equatorial and axial positions, respectively. In the solid state structures of compounds 3 and 4 both nitrogen atoms of the bipy ligand occupy equatorial positions. The NMR spectroscopy reveals a fluxional behavior of compounds 1, 2, 5, and 6 in solution. Although the fluxional behavior of compounds 5 and 6 ceased at about −40 °C, that of compound 1 could not be stopped even at −90 °C. Their low temperature conformations are consistent with their solid state structures. Both the endo- and exo-isomers coexist in solution for compounds 3 and 4.  相似文献   

15.
Reaction of [Mn2(CO)9(NCMe)] with tetrahydropyrimidine-2-thione (thpymSH) at 25 °C furnishes the mono- and dinuclear complexes [Mn(CO)411-SCNHC3H6NCO)] (2) and [Mn2(CO)6(μ-thpymS)2] (1), respectively. Carbon-nitrogen coupling is observed in compound 2 resulting in the formation of κ11-SCNHC3H6NCO ligand while compound 1 adopts a centrosymmetric structure. Reaction of 1 with [Os3(CO)10(NCMe)2] at 80 °C affords the mixed Mn-Os cluster [MnOs3(CO)133-thpymS)] (3) which possesses a butterfly skeleton of four metal atoms whereas with Ru3(CO)12 at 110 °C gives the mixed Mn-Ru complex [MnRu3(CO)144-S)(κ11-thpym)] (4). In contrast, treatment of 1 with Fe3(CO)12 at 80 °C furnishes two triiron complexes [Fe3(CO)93-S)(μ311-C4H6N2)] (5) and [Fe3(CO)83-S)21-C4H8N2)] (6). The former also results from the direct reaction of thpymSH with Fe3(CO)12 and reacts with H2S to afford 6. The molecular structures of all these new complexes have been determined by X-ray diffraction studies.  相似文献   

16.
17.
Cationic metal complexes of dipicolinic acid (dipicH2) are stabilized by [Ce(dipic)3]2− ions in the three isomorphous crystals [M(dipicH2)(OH2)3][Ce(dipic)3] · 3H2O (M = Ni, 1; Cu, 2; Zn, 3). Magnetic dilution provided by the bulky anions leads to well-resolved EPR spectra in polycrystalline samples of 2. The cations have 4+2 coordination, the carbonyl atom of the carboxylic acid groups coordinating weakly from trans positions. In the case of 2 this steric distortion is augmented by Jahn–Teller distortion. All the three structures are satisfactorily modelled by calculations based on density functional theory (DFT). The switch of the Jahn–Teller axis upon deprotonation of the complex, leading to the neutral species Cu(dipic)(H2O)3, is also reproduced by DFT. Electronic transition energies as well as the g-tensor component of the d9 complex obtained are in good agreement with experiment. However, the calculated hyperfine coupling constants are in error. DFT also fails to satisfactorily account for the electronic transition in the d8 ion in 1.  相似文献   

18.
The dipalladium complexes, [PdCl(μ-MeN{P(OR)2}2)]2 (R = CH2CF3, 1a; Ph, 1b) react with [Mo25-C5H5)2(CO)6] in boiling benzene to afford the molybdenum-palladium heterometallic complexes, [(η5-C5H5)(CO)Mo(μ-MeN{P(OR)2}2)2PdCl] (R = CH2CF3, 3a; Ph, 3b), [(η5-C5H5)Mo(μ3-CO)2(μ-MeN{P(OR)2}2)2Pd2Cl], (R = CH2CF3, 5a; Ph, 5b), [(η5-C5H5)(Cl)Mo(μ2-CO)(μ2-Cl)(μ-MeN{P(OR)2}2)PdCl], (R = CH2CF3, 6a; Ph, 6b) and also the mononuclear complex [Mo(CO)Cl(η5-C5H5)(κ2-MeN{P(OR)2}2)], (R = Ph, 4b). These complexes have been separated by column chromatography and are characterised by elemental analysis, IR, 1H, 31P{1H} NMR data. The structures of 1a, 3a, 4b, 5b and 6a have been confirmed by single crystal X-ray diffraction. The CO ligands in 5b and 6a adopt a semi-bridging mode of bonding; the Mo-CO distances (1.95-1.97 Å) are shorter than the Pd-CO distances (2.40-2.48 Å). The Pd-Mo distances fall in the range, 2.63-2.86 Å. The reaction of [Mo25-C5H5)2(CO)6] with MeN{P(OPh)2}2 in toluene gives [Mo2(CO)45-C5H5)21-MeN{P(OPh)2}2)2] (2) in which the diphosphazane acts as a monodentate ligand.  相似文献   

19.
Treatment of (C5H4SiMe2tBu)2LnR with 1 equiv of elemental sulfur in toluene at ambient temperature gives dimeric complexes [(C5H4SiMe2tBu)2Ln(μ-SR)]2 [R = Me, Ln = Yb (1), Er (2), Dy (3), Y (4); R = nBu, Ln = Yb (5), Dy (6)]. All these complexes have been characterized by elemental analysis, IR and mass spectroscopies. The structures of complexes 1, 3, 5 and 6 are also determined through X-ray single crystal diffraction analysis, indicating that only one sulfur atom from elemental sulfur inserts into Ln–C σ-bond.  相似文献   

20.
The first cationic samarium phenoxide complex, [(ArO)2Sm(DME)2][BPh4] · THF (ArO = 2,6-di-tert-butyl-4-metyl-phenoxide) (1), has been synthesized by one-electron oxidation reaction of (ArO)2Sm(THF)3 with AgBPh4 in high yield and structurally characterized. The complex 1 can be used as a single-component catalyst for the ring-opening polymerization of ε-caprolactone (ε-CL) with high activity. The activity of the complex 1 is much higher than that of the parent neutral complex (ArO)3Sm(THF)2, and is comparable to that of the divalent complex (ArO)2Sm(THF)3. A coordination-insertion polymerization mechanism was supposed according to the end-group 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