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 224 毫秒
1.
Imidazolium salts bearing triazole groups are synthesized via a copper catalyzed click reaction, and the silver, palladium, and platinum complexes of their N‐heterocyclic carbenes are studied. [Ag4(L1)4](PF6)4, [Pd(L1)Cl](PF6), [Pt(L1)Cl](PF6) (L1=3‐((1‐benzyl‐1H‐1,2,3‐triazol‐4‐yl)methyl)‐1‐(pyrimidin‐2‐yl)‐1H‐imidazolylidene), [Pd2(L2)2Cl2](PF6)2, and [Pd(L2)2](PF6)2 (L2=1‐butyl‐3‐((1‐(pyridin‐2‐yl)‐1H‐1,2,3‐triazol‐4‐yl)methyl)imidazolylidene) have been synthesized and fully characterized by NMR, elemental analysis, and X‐ray crystallography. The silver complex [Ag4(L1)4](PF6)4 consists of a Ag4 zigzag chain. The complexes [Pd(L1)Cl](PF6) and [Pt(L1)Cl](PF6), containing a nonsymmetrical NCN ′ pincer ligand, are square planar with a chloride trans to the carbene donor. [Pd2(L2)2Cl2](PF6)2 consists of two palladium centers with CN2Cl coordination mode, whereas the palladium in [Pd(L2)2](PF6)2 is surrounded by two carbene and two triazole groups with two uncoordinated pyridines. The palladium compounds are highly active for Suzuki–Miyaura cross coupling reactions of aryl bromides and 1,1‐dibromo‐1‐alkenes in neat water under an air atmosphere.  相似文献   

2.
Co-condensation of atoms of Re, Ru, Rh, Ir and Pt with oxalyl chloride gives metal chloro-carbonyl derivatives which may be used as precursors to the compounds [Re(CO)4Cl]2, [Ru(PMe3)3(CO)Cl2], α-[Ru(CO)3Cl(μ-Cl)]2, [Ru(PPh3)2(CO)2Cl2], [Rh(CO)2(μ-Cl)]2, [Rh(PPh3)2COCl], [Ir(PPh3)(CO)2Cl3] and cis-Pt(CO)2Cl2. Molybdenum atoms with oxalyl chloride give molybdenum-chloro derivatives.  相似文献   

3.
The title complex, μ‐octane‐1,8‐dioato‐bis[bis(3‐aminopyridine)chloro(methanol)cobalt(II)], [Co2(C8H12O4)Cl2(C5H6N2)4(CH4O)2], is located on a crystallographic centre of inversion. The coordination around each of the Co centres is distorted octa­hedral, involving two N, three O and one Cl atom. Discrete dimers are connected in a three‐dimensional arrangement through N—H⋯O, N—H⋯Cl and O—H⋯O hydrogen‐bond inter­actions.  相似文献   

4.
The reductive carbonylation of silica-supported OsCl3·3H2O was investigated under atmospheric pressure of CO or of a mixture of CO and H2O at relatively mild temperatures. Starting from OsCl3·3H2O a mixture of physisorbed α-[Os(CO)3Cl2]2, cis-[Os(CO)4Cl2] and a species bound to the surface silanol groups, [Os(CO)3Cl2(HOSi)], is formed working at 100°C under CO. At higher temperatures [Os(CO)3Cl2(HOSi)] is the major surface species. Attempts to reduce further on the surface these OsII chlorocarbonyl species failed due to their easy sublimation and to the difficult removal of the chloro ligands. However, when the silica is treated with a weak base such as NaHCO3, α- or β-[Os(CO)3Cl2]2 supported on silica may be reduced with CO to [Os3(CO)12], with CO and H2O to a mixture of [OS3(CO)12] and [H4Os4(CO)12], and finally with H2 to [H4Os4(CO)12]. These reduction processes occur via the anchored [Os(CO)3(OSi)2]n species, as intermediate surface species.  相似文献   

5.
The tri­chloro‐bridged dinuclear RuII complex tri‐μ‐chloro‐bis{[1,1,1‐tris­(di­phenyl­phosphino­methyl)­ethane‐κ3P,P′,P′′]ruthenium(II)} hexa­fluoro­phosphate ethanol solvate, [Ru2Cl3(tripod)2]PF6·C2H6O, containing the tripod [1,1,1‐tris­(di­phenyl­phosphino­methyl)­ethane, C41H39P3] ligand, was unexpectedly obtained from the reaction of [RuIIICl3(tripod)] with 1,4‐bis­(di­phenyl­phosphino)­butane (dppb), followed by pre­cipitation with NH4PF6. The magnetic moment of the compound at room temperature indicates that the dinuclear [Ru2(μ‐Cl)3(tripod)2]+ cation is diamagnetic. A single‐crystal X‐ray structure determination revealed that the two Ru atoms are bridged by the three Cl atoms. The coordination sphere of each Ru atom is completed by the three P atoms of a tripod ligand. The two P3Ru units are exactly eclipsed, while the bridging Cl atoms are staggered with respect to the six P atoms. The Ru⋯Ru distance is 3.3997 (7) Å and the mean Cl—Ru—Cl bond angle is 77.7°.  相似文献   

6.
The two title dinuclear copper(II) complexes, [Cu2Cl4(C17H20Cl2N2)2], (I), and [Cu2Cl4(C19H22N2O4)2], (II), have similar coordination environments. In each complex, the asymmetric unit consists of one half‐molecule and the two copper centres are bridged by a pair of Cl atoms, resulting in complexes with centrosymmetric structures containing Cu(μ‐Cl)2Cu parallelogram cores; the Cu...Cu separations and Cu—Cl—Cu angles are 3.4285 (8) Å and 83.36 (3)°, respectively, for (I), and 3.565 (2) Å and 84.39 (7)° for (II). Each Cu atom is five‐coordinated and the coordination geometry around the Cu atom is best described as a distorted square‐pyramid with a τ value of 0.155 (3) for (I) and 0.092 (7) for (II). The apical Cu—Cl bond length is 2.852 (1) Å for (I) and 2.971 (2) Å for (II). The basal Cu—Cl and Cu—N average bonds lengths are 2.2673 (9) and 2.030 (2) Å, respectively, for (I), and 2.280 (2) and 2.038 (6) Å for (II). The molecules of (I) are linked by one C—H...Cl hydrogen bond into a complex [10] sheet. The molecules of (II) are linked by one C—H...Cl and one N—H...O hydrogen bond into a complex [100] sheet.  相似文献   

7.
The mononuclear title complex, [MnCl2(C3H4N2)2(H2O)2], is located on a crystallographic inversion center. The MnII ion is coordinated by two imidazole ligands [Mn—N = 2.2080 (9) Å], two Cl atoms [Mn—Cl = 2.5747 (3) Å] and two water molecules [Mn—O = 2.2064 (8) Å]. These six monodentate ligands define an octahedron with almost ideal angles: the adjacent N—Mn—O, N—Mn—Cl and O—Mn—Cl angles are 90.56 (3), 92.04 (2) and 90.21 (2)°, respectively. Hydrogen bonds between the coordinated water molecules and Cl atoms form a two‐dimensional network parallel to (100) involving R42(8) rings. The two‐dimensional networks link into a three‐dimensional framework through weaker N—H...Cl interactions. Thermogravimetric analysis results are in accordance with the water‐coordinated character of the substance and its dehydration in two successive steps.  相似文献   

8.
The first selenite chloride hydrates, Co(HSeO3)Cl · 3 H2O and Cu(HSeO3)Cl · 2 H2O, have been prepared from solution and characterised by single‐crystal X‐ray diffraction. The cobalt phase adopts an unusual “one‐dimensional” structure built up from vertex‐sharing pyramidal [HSeO3]2–, and octahedral [CoO2(H2O)4]2– and [CoO2(H2O)2Cl2]4– units. Inter‐chain bonding is by way of hydrogen bonds or van der Waals' interactions. The atomic arrangement of the copper phase involves [HSeO3]2– pyramids and Jahn‐Teller distorted [CuCl2(H2O)4] and [CuO4Cl2]8– octahedra, sharing vertices by way of Cu–O–Se and Cu–Cl–Cu bonds. Crystal data: Co(HSeO3)Cl · 3 H2O, Mr = 276.40, triclinic, space group P 1 (No. 2), a = 7.1657(5) Å, b = 7.3714(5) Å, c = 7.7064(5) Å, α = 64.934(1)°, β = 68.894(1)°, γ = 71.795(1)°, V = 337.78(7) Å3, Z = 2, R(F) = 0.036, wR(F) = 0.049. Cu(HSeO3)Cl · 2 H2O, Mr = 263.00, orthorhombic, space group Pnma (No. 62), a = 9.1488(3) Å, b = 17.8351(7) Å, c = 7.2293(3) Å, V = 1179.6(2) Å3, Z = 8, R(F) = 0.021, wR(F) = 0.024.  相似文献   

9.
Synthetic procedures are described that allow access to the new complexes cis-[Mo2O5(apc)2], cis-[WO2(apc)2], trans-[UO2(apc)2], [Ru(apc)2(H2O)2], [Ru(PPh3)2(apc)2], [Rh(apc)3], [Rh(PPh3)2(apc)2]ClO4, [M(apc)2], [M(PPh3)2(apc)]Cl, [M(bpy)(apc)]Cl (M(II) = Pd, Pt), [Pd(bpy)(apc)Cl], [Ag(apc)(H2O)2] and [Ir(bpy)(Hapc)2]Cl3, where Hapc, is 3-aminopyrazine-2-carboxylic acid. These complexes were characterized by physico-chemical and spectroscopic techniques. Both Hapc and several of its complexes display significant anticancer activity against Ehrlich ascites tumour cells (EAC) in albino mice.  相似文献   

10.
The accelerated formation of 2,3-diphenylquinoxalines in microdroplets generated in a nebulizer has been investigated by competition experiments in which equimolar quantities of 1,2-phenylenediamine, C6H4(NH2)2, and a 4-substituted homologue, XC6H3(NH2)2 [X = F, Cl, Br, CH3, CH3O, CO2CH3, CF3, CN or NO2], or a 4,5-disubstituted homologue, X2C6H2(NH2)2 [X = F, Cl, Br, or CH3], compete to condense with benzil, (C6H5CO)2. Electron-donating substituents (X = CH3 and CH3O) accelerate the reaction; in contrast, electron-attracting substituents (X = F, Cl, Br and particularly CO2CH3, CN, CF3 and NO2) retard it. A structure–reactivity relationship in the form of a Hammett correlation has been found by analyzing the ratio of 2,3-diphenylquinoxaline and the corresponding substituted-2,3-diphenylquinoxaline, giving a ρ value of −0.96, thus confirming that the electron density in the aromatic ring of the phenylenediamine component is reduced in the rate-limiting step in this accelerated condensation. This correlation shows that the phenylenediamine acts as a nucleophile in the reaction.  相似文献   

11.
Cobalt(II), nickel(II), and copper(II) complexes containing 5,12-di(4-bromophenyl)-7,14-dimethyl-1,2,4,8,9,11-hexaazacyclotetradeca-7,14-diene-3,10-dione (H2L1) and 5,12-diphenyl-7,14-dimethyl-1,2,4,8,9,11-hexaazacyclotetradeca-7,14-diene-3,10-dione (H2L2) have been synthesized. All complexes were characterized by elemental analysis, MALDI TOF-MS spectrometry, and electronic absorption spectroscopy. The crystal structures of two compounds, [Cu2(H2L1)Cl4]n and [NiL2], were determined by X-ray powder diffraction. In the polymeric [Cu2(H2L1)Cl4]n, the Cu2Cl4 units and H2L1 molecules are situated on inversion centers. Each Cu(II) has a distorted trigonal-bipyramidal coordination environment formed by N and O from H2L1 [Cu–N 2.340(14)?Å, Cu–O 1.952(11)?Å], two bridging chlorides [Cu–Cl 2.332(5), 2.279(5)?Å] and one terminal chloride [Cu–Cl 2.320(6)?Å]. In the [NiL2] complex, the Ni(II) situated on inversion center has a distorted square-planar coordination environment formed by four nitrogens from L2 [Ni–N 1.860(11), 1.900(11)?Å].  相似文献   

12.
The title compound, [NiCl(C12H32N6)(H2O)]Cl3·3H2O, has the bis­(diamine)‐substituted cyclic tetra­amine in a planar coordination to triplet ground‐state NiII [average Ni—N = 2.068 (3) Å], with a chloride ion [Ni—Cl = 2.4520 (5) Å] and a water mol­ecule [Ni—O = 2.177 (2) Å] coordinated in the axial sites. The amine substituents are protonated and equatorially oriented. The amine groups, ammonium groups, water molecules and chloride ions are linked by an extensive hydrogen‐bonding network.  相似文献   

13.
《Polyhedron》1987,6(5):1009-1015
Reactions of 2-mercapto-3-phenyl-4-quinazolinone (LH) with RuCl3·xH2O and RhCl3·xH2O afforded the compounds [RuL2Cl(H2O)]H2O, [RuL2Cl·DMFI and RhL(LH)Cl2·2H2O. Reactions of LH with RuCl3·xH2O in the presence of N-heterocyclic bases led to the formation of complexes of type [RuL2ClB]·H2O (B = pyridine, 3-picoline or imidazole) and [RuLCl2(o-phen)] H2O (o-phen = 1, 10-phenanthroline). These complexes were characterized on the basis of analytical, conductivity, magnetic, IR and electronic spectral and ESR studies. Tentative structures for the complexes are proposed.  相似文献   

14.
The influence of the potentially chelating imino group of imine‐functionalized Ir and Rh imidazole complexes on the formation of functionalized protic N‐heterocyclic carbene (pNHC) complexes by tautomerization/metallotropism sequences was investigated. Chloride abstraction in [Ir(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 a ) (cod=1,5‐cyclooctadiene, Dipp=2,6‐diisopropylphenyl) with TlPF6 gave [Ir(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 a +[PF6]?). Plausible mechanisms for the tautomerization of complex 1 a to 3 a +[PF6]? involving C2?H bond activation either in 1 a or in [Ir(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 a +[PF6]?) were postulated. Addition of PR3 to complex 3 a +[PF6]? afforded the eighteen‐valence‐electron complexes [Ir(cod)(PR3){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 7 a +[PF6]? (R=Ph) and 7 b +[PF6]? (R=Me)). In contrast to Ir, chloride abstraction from [Rh(cod)Cl{C3H3N2(DippN=CMe)‐κN3}] ( 1 b ) at room temperature afforded [Rh(cod){C3H3N2(DippN=CMe)‐κN3}2]+[PF6]? ( 6 b +[PF6]?) and [Rh(cod){C3H3N2(DippN=CMe)‐κ2(C2,Nimine)}]+[PF6]? ( 3 b +[PF6]?) (minor); the reaction yielded exclusively the latter product in toluene at 110 °C. Double metallation of the azole ring (at both the C2 and the N3 atom) was also achieved: [Ir2(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 10 ) and the heterodinuclear complex [IrRh(cod)2Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 12 ) were fully characterized. The structures of complexes 1 b , 3 b +[PF6]?, 6 a +[PF6]?, 7 a +[PF6]?, [Ir(cod){C3HN2(DippN=CMe)(DippN=CH)(Me)‐κ2(N3,Nimine)}]+[PF6]? ( 9 +[PF6]?), 10? Et2O ? toluene, [Ir2(CO)4Cl{μ‐C3H2N2(DippN=CMe)‐κ2(C2,Nimine),κN3}] ( 11 ), and 12? 2 THF were determined by X‐ray diffraction.  相似文献   

15.
The title compound, [Sn4(CH3)8(C13H8Cl2NO2)2(C2H5O)2O2], is a centrosymmetric dimer, with three linearly fused four‐membered Sn—O—Sn—O rings. The coordination poly­hedron of the Sn atom bonded to the carboxyl­ate can be described as trigonal–bipyramidal distorted toward square‐pyramidal. That of the second Sn atom is similar, but the distortion towards square‐pyramidal geometry is greater. The Sn—O and Sn—C distances are 2.020 (2)–2.226 (2) and 2.096 (4)–2.114 (4) Å, respectively. The benzene rings of the 2‐[(2,3‐dichloro­phenyl)­amino]benzoate ligand subtend an angle of 50.49 (17)°; the conformation of the ligand is stabilized by intra­molecular N—H⋯Cl and N—H⋯O hydrogen bonds. The structure is assembled viaπ–π stacking inter­actions to form chains parallel to [10].  相似文献   

16.
Four NNN tridentate ligands L1–L4 containing 2‐methoxypyridylmethene or 2‐hydroxypyridylmethene fragment were synthesized and introduced to ruthenium centers. When (HOC5H3NCH2C5H3NC5H7N2) (L2) and (HOC5H3NCH2C5H3NC6H6N3) (L4) reacted with RuCl2(PPh3)3, two ruthenium chloride products Ru(L2)(PPh3)Cl2 ( 1 ) and Ru(L4)(PPh3)Cl2 ( 2 ) were isolated, respectively. Reactions of (MeOC5H3NCH2C5H3NC5H7N2) (L1) and (MeOC5H3NCH2C5H3NC6H6N3) (L3) with RuCl2(PPh3)3 in the presence of NH4PF6 generated two dicationic complexes [Ru(L1)2][PF6]2 ( 3 ) and [Ru(L3)2][PF6]2 ( 4 ), respectively. Complex 1 reacted with CO to afford product [Ru(L2)(PPh3)(CO)Cl][Cl]. The catalytic activity for transfer hydrogenation of ketones was investigated. Complex 1 showed the highest activity, with a turnover frequency value of 1.44 × 103 h?1 for acetophenone, while complexes 3 and 4 were not active.  相似文献   

17.
Interaction of copper(II) salts with 2,2′-dipyridylamine (1), N-cyclohexylmethyl-2,2′-dipyridylamine (2), di-2-pyridylaminomethylbenzene (3), 1,2-bis(di-2-pyridylaminomethyl)-benzene (4), 1,3-bis(di-2-pyridylaminomethyl)benzene (5), 1,4-bis(di-2-pyridylaminomethyl)benzene (6), 1,3,5-tris(di-2-pyridylaminomethyl)benzene (7) and 1,2,4,5-tetrakis(di-2-pyridylaminomethyl)benzene (8) has yielded the following complexes: [Cu(2)(μ-Cl)Cl]2, [Cu(3)(μ-Cl)Cl]2 · H2O, [Cu2(4)(NO3)4], [Cu2(5)(NO3)4] · 2CH3OH, [Cu2(6)(CH3OH)2(NO3)4], [Cu4(8)](NO3)4] · 4H2O while complexation of palladium(II) with 1, 4, 5 and 6 gave [Pd(1)2](PF6)2 · 2CH3OH, [Pd2(4)Cl4], [Pd2(4)(OAc)4], [Pd2(5)Cl4], [Pd2(6)Cl4] and [Pd2(6)(OAc)4] · CH2Cl2, respectively. X-ray structures of [Cu(2)(μ-Cl)Cl]2, [Cu(3)(μ-Cl)Cl]2 · 2C2H5OH, [Cu2(6)(CH3OH)2(NO3)4], [Pd(1)2](PF6)2 · 2CH3OH, [Pd2(4)(OAc)4] · 4H2O and [Pd2(6)(OAc)4] · 2CH2Cl2 are reported. In part, the inherent flexibility of the respective ligands has resulted in the adoption of a diverse range of coordination geometries and lattice arrangements, with the structures of [Pd2(4)(OAc)4· 4H2O and [Pd2(6)(OAc)4] · 2CH2Cl2, incorporating the isomeric ligands 4 and 6, showing some common features. Liquid–liquid (H2O/CHCl3) extraction experiments involving copper(II) and 13, 5, 7and 8 show that the degree of extraction depends markedly on the number of dpa-subunits (and concomitant lipophilicity) of the ligand employed with the tetrakis-dpa derivative 8 acting as the most efficient extractant of the six ligand systems investigated.  相似文献   

18.
The reactions of AuIII, PtII and PdII complexes with 2-pyridinecarboxaldehyde (2CHO-py) have been examined in protic (H2O, MeOH, EtOH) and aprotic (DMF, CH2Cl2) solvents. Compounds in which the pyridine ligand is N-coordinated, either in the original aldehydic form or in a new form derived from addition of one or two protic molecules, have been isolated, namely: [Au(2CHO-py · H2O)Cl3], [Au(2CHO-py · MeOH)Cl3], [Au(2CHO-py · 2EtOH)Cl3], cis-[Pt(2CHO-py)2Cl2], trans-[Pd(2CHO-py)2Cl2], trans-[Pt(dmso)(2CHO-py)Cl2], [Pt{C5H4N-(CH2SMe)}Cl(2CHO-py)](ClO4), [Pt(terpy)(2CHOpy)](ClO4)2, [Pt(terpy)(2CHO-py · H2O)](ClO4)2 (terpy = 2,2′:6′,2′′-terpyridine). 1H-n.m.r. experiments show that the addition of the protic molecule(s) to the PtII and PdII complexes is reversible. The effects of the nature of the metal ion and the ancillary ligands as well as of the total charge of the complexes on the relative stability of the addition products are discussed. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

19.
The thermal reaction of Ru3(CO)12 with ethacrynic acid, 4‐[bis(2‐chlorethyl)amino]benzenebutanoic acid (chlorambucil), or 4‐phenylbutyric acid in refluxing solvents, followed by addition of two‐electron donor ligands (L), gives the diruthenium complexes Ru2(CO)4(O2CR)2L2 ( 1 : R = CH2O‐C6H2Cl2‐COC(CH2)C2H5, L = C5H5N; 2 : R = CH2O‐C6H2Cl2‐COC(CH2)C2H5, L = PPh3; 3 : R = C3H6‐C6H4‐N(C2H4‐Cl)2, L = C5H5N; 4 : R = C3H6‐C6H4‐N(C2H4‐Cl)2, L = PPh3; 5 : R = C3H6‐C6H5, L = C5H5N; 6 : R = C3H6‐C6H5, L = PPh3). The single‐crystal structure analyses of 2 , 3 , 5 and 6 reveal a dinuclear Ru2(CO)4 sawhorse structure, the diruthenium backbone being bridged by the carboxylato ligands, while the two L ligands occupy the axial positions of the diruthenium unit.  相似文献   

20.
Some new complexes of rhenium containing monocoordinating acetylacetone viz. [Re(NO)(acac)2(Cl)H2O], Re(NO)(acac)3(acacH)(H2O)2 have been synthesized by reacting acetylacetone with either Re(NO)Cl3 · H2O or Re(NO)(OH)3 · H2O. The preparation of [Re(NO)(py)(acac)2(Cl) · H2O], [Re(NO)(acac)(OH)(Cl) · H2O], [Re2(NO)2(acacH)3(acac)2 · Cl4], [Re(NO)(acac)2(OH) · H2O] · 2 H2O have also been described. These complexes were characterized through their elemental analyses, u.v., vis, i.r., 1Hn.m.r., magnetic, and conductance data.  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