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.
The diamagnetic complexes [Pd2(H2L1)Cl4] (I), [Pd2(H2L2)Cl4] (II), and Pd2(H2L3)Cl4(III) with chiral ligands derived from the natural monoterpenoid (R)-(+)-limonene are obtained (H2 L1 is ethylenediamine dioxime, H2L2 is piperazine dioxime, and H2L3 is propylenediamine dioxime). According to X-ray diffraction data, the crystal structures of complexes I and II are composed of binuclear acentric molecules. The coordination polyhedra PdN2Cl2 are trapeziums (squares distorted in a tetrahedral manner) made up of two N atoms of the tetradentate bridging cyclic ligands H2L1 and H2L2 and two Cl atoms. The fragments PdCl2 are trans in the complexes. The 13C and 1H NMR spectra of complexes I and II in CDCl3 also suggest their binuclear structures.  相似文献   

2.
This research was an outgrowth of previous reactions with [Pd13Ni13(CO)34]4? which produced a tetragonal crystal form of Pd23(CO)20(PEt3)10 (1) that has the same cuboctahedral-based Pd23 framework with an identical number of PEt3 ligands but two fewer CO ligands than the monoclinic crystal form of Pd23(CO)22(PEt3)10 (3) originally reported from reactions with Pd10(CO)12(PEt3)6. A subsequent investigation presented herein to establish whether the carbonyl capacity is influenced by the nature of the phosphine ligands has led to syntheses of Pd23(CO) x (PR3)10 [R3=Et3 (1), Bu n 3 (4), and Me2Ph (5)] with 20 CO ligands (x=20) from corresponding Pd10(CO)12(PR3)6 precursors either by deligation with Pd(OAc)2, CF3CO2H, Ni(1,5-COD)2, [NMe4]2[Ni6(CO)12], or HCO2H or by spontaneous enlargement; yields varied from 15 to 79%. Although attempts to obtain the original Pd23(CO)22(PEt3)10 (3) were unsuccessful, a highly significant outcome was the isolation (one time) of another monoclinic crystal form possessing the triethylphosphine Pd23(CO) x (PEt3)10 cluster with 21 COs (2). Both the compositions and atomic arrangements for each of five Pd23 clusters [1a (solvated); 1b (unsolvated); 2, 4, and 5] were unambiguously established from low-temperature single-crystal CCD X-ray crystallographic determinations in accordance with their nearly identical IR carbonyl frequencies. Solution 31P{1H} NMR spectra of 1 and 4 at room temperature displayed three distinct signals with expected integral ratios of 2/4/4 that are consistent with the solid-state structures of Pd23(CO)20(PR3)10 [R3=Et3 (1), Bu n 3 (4)] remaining intact in solution. The metal-core geometries of all of these Pd23(CO) x (PR3)10 clusters, including the thermodynamically stable ones with 20 CO ligands and the kinetic products with additional CO ligands (x=21, 22), are essentially the same. The common Pd23 core may be best described as possessing a centered hexacapped cuboctahedral Pd19 kernel (alternatively denoted as a centered ν2 Pd19 octahedron) with four edge-connected exopolyhedral wingtip Pd(exo) atoms that reduce the pseudo metal-core symmetry from Oh to D2h. The 10 PR3 ligands are linked to the six tetracapped Pd(cap) and four edge-capped wingtip Pd(exo) atoms; the latter four Pd(exo) atoms are each composed of four trigonal-planar Pd(μ2-CO)2(PR3) units. These crystallographic results provide compelling geometrical evidence for a heretofore unknown stereochemical example involving variable carbonyl ligation (x=20, 21, 22) of a close-packed nanosized Pd n (CO) x (PR3) y cluster (in this case with identical PEt3 ligands) without significant changes being induced in either the overall metal-core architecture or steric dispositions of the same number of PR3 ligands. These experimental findings have particular relevance to the long-standing Muetterties cluster/surface science analogy in showing that the different number (as well as different modes) of carbonyl ligations observed in these large metal carbonyl clusters are directly related to pressure-induced dissociative/nondissociative migratory coverages in CO chemisorptions on metal surfaces. The observed expanded capacity of CO coordination on the same Pd23 polyhedron without notable changes in geometry is no doubt a consequence of its virtually nanosized metal-core architecture; distances between outermost centrosymmetrically related pairs of Pd(cap) and Pd(exo) atoms in the Pd23 framework are 0.8 and 0.9 nm, respectively. An electrochemical (CV) study revealed that 1 undergoes one quasi-reversible two-electron reduction to 1 2? (E1/2=?0.91 V) and two consecutive quasi-reversible one-electron oxidations to 1/1 + at E1/2=0.08 V and 1 +/1 2+ at E1/2=0.32 V (THF; Ag/AgCl as reference electrode). A stereochemical/electronic analysis with the isostructural Au2Pd21(CO)20(PEt3)10 analogue (9) and resulting implications are given.  相似文献   

3.
吕兴强  卢荣 《高分子科学》2014,32(6):768-777
From the self-assembly of the typical Salen-type Schiff-base ligand H2L and Zn(OAc)2·2H2O in the molar ratio of 1:1 or 1:2, the mononuclear [Zn(L)(H2O)](1) or binuclear [Zn2(L)(OAc)2(H2O)](2) are obtained, respectively. For both complexes 1 and 2, the unsaturated five-coordinate coordination environment to the catalytic active centers(Zn2+ ions) permits the monomer insertion for the effective solution copolymerization of cyclohexene oxide and maleic anhydride. All the solution copolymerizations afford poly(ester-co-ether)s, while lower catalyst and co-catalyst concentrations are helpful for the formation of alternating polyester. Of the three co-catalysts, 4-(dimethylamino)pyridine is found to be the most efficient, while an excess thereof is detrimental for chain growth of the copolymers.  相似文献   

4.
Heating of the compounds (RC5H4)Fe(CO)2TePh (R = H (I) and Me (II)) in heptane afforded the dinuclear complexes [(RC5H4)Fe(CO)TePh]2 (III and IV, respectively). By oxidation with Fc+PF 6 ? , these complexes were transformed into the paramagnetic cationic complexes [(RC5H4)Fe(CO)TePh]2PF6 (V and VI, respectively). Structures III–V and [(C5H5)Fe(CO)SPh]2PF6 (VII) were characterized by X-ray diffraction.  相似文献   

5.
[Cp4Fe4(CO)4] (1) reacts with p-BrC6H4Li and MeOH in sequence to afford the functionalized cluster [Cp3Fe4(CO)4(C5H4-p-C6H4Br)] (2), while the reaction of 2 with n-BuLi and MeOH produces [Cp2Fe4(CO)4(C5H4Bu)(C5H4-p-C6H4Br)] (3). The double cluster [Cp3Fe4(CO)4(C5H4)]2(p-C6H4) (4) has been prepared by treatment of [Cp4Fe4(CO)4] with p-C6H4Li2 and MeOH in sequence. The electrochemistry of 2 and 4, as well as the crystal structure of 4 have been investigated.  相似文献   

6.
We report a phase diagram (on the mole fraction scale) for the [Th(NO3)4(TBP)2]-decane-[UO2(NO3)2(TBP)2](1-2-3) ternary liquid system, where TBP stands for tributyl phosphate, at T = 298.15 K. This system is characterized by a homogeneous solution field and a two-liquid field (immiscibility field); one phase (phase I) is enriched in [Th(NO3)4(TBP)2] and [UO2(NO3)2(TBP)2], and the other (phase II) is enriched in decane. Molecular interaction parameters and excess Gibbs energies G ex were calculated for the binary systems and the ternary liquid system along the binodal curve proceeding from miscibility in the [Th(NO3)4(TBP)2]-decane system and the ternary system and using equations of the NTRL model. For the ternary system, G ex > 0. G ex decreases in the following order of pairs of liquids: (1, 2) > (2, 3) > (1, 3).  相似文献   

7.
Four new butterfly Fe/S cluster complexes bearing 2,6-(CH2)2C5H3N or (CH2)2 groups, as the active site models of [FeFe]-hydrogenase, have been prepared by condensation reaction and structurally characterized. Treatments of the parent complex Fe2(CO)6[(μ-SCH2)2CHCO2H] (A) with 2,6-(HOCH2)2C5H3N or HOCH2CH2OH in the presence of 4-dimethylaminopyridine and dicyclohexylcarbodiimide afforded the single-butterfly Fe/S complexes Fe2(CO)6[(μ-SCH2)2CHC(O)OCH2(2,6-C5H3N)CH2OH] (1) and Fe2(CO)6[(μ-SCH2)2CHC(O)OCH2CH2OH] (3) and the double-butterfly Fe/S complexes [Fe2(CO)6(μ-SCH2)2CHC(O)OCH2]2(2,6-C5H3N) (2) and [Fe2(CO)6(μ-SCH2)2CHC(O)OCH2]2 (4). The new complexes 14 were fully characterized by elemental analysis, ESI-MS, IR, and 1H (13C) NMR spectroscopy.  相似文献   

8.
The zirconium nitrate complexes (NO2)[Zr(NO3)3(H2O)3]2(NO3)3 (1), Cs[Zr(NO3)5] ((2), (NH4)[Zr(NO3)5](HNO3) (3), and (NO2)0.23(NO)0.77[Zr(NO3)5] ((4) were prepared by crystallization from nitric acid solutions in the presence of H2SO4 or P2O5. The complexes were characterized by X-ray diffraction. The crystal structure of 1 consists of nitrate anions, nitronium cations, and [Zr(NO3)3(H2O)3]+ complex cations in which the ZrIV atom is coordinated by three water molecules and three bidentate nitrate groups. The coordination polyhedron of the ZrIV atom is a tricapped trigonal prism formed by nine oxygen atoms. The island structures of 2 and 3 contain [Zr(NO3)5]? anions and Cs+ or NH4 + cations, respectively. In addition, complex 3 contains HNO3 molecules. Complex 4 differs from (NO2)[Zr(NO3)5] in that three-fourth of the nitronium cations in 4 are replaced by nitrosonium cations NO+, resulting in a decrease in the unit cell parameters. In the [Zr(NO3)5]? anion involved in complexes 2–4, the ZrIV atom is coordinated by five bidentate nitrate groups and has an unusually high coordination number of 10. The coordination polyhedron is a bicapped square antiprism.  相似文献   

9.
The synthesis, derivatization and coordination behavior of a new aminobis(diphosphonite), PhN{P(OC6H4OMe-o)2}2 (1) is described. The ligand 1 reacts with H2O2, elemental sulfur or selenium to give the corresponding dichalcogenides PhN{P(E)(OC6H4OMe-o)2}2 (E = O, 2; S, 3; Se, 4) in good yield. Reactions of 1 with Mo(CO)6, Pd(NCCH3)2Cl2 and Pt(COD)Cl2 resulted in the formation of the chelate complexes, Mo(CO)4[PhN{P(OC6H4OMe-o)2}2] (5) and MCl2[PhN{P(OC6H4OMe-o)2}2] (M = Pd,7; M = Pt, 8) whereas in the reaction of 1 with [CpFe(CO)2]2, one of the P-N bonds cleaves due to the metal assisted hydrolysis to give a mononuclear complex, [CpFe(CO){P(O)(OC6H4OMe-o)2}{PhN(H)(P(OC6H4OMe-o)2)}] (6). The molecular structures of 1, 4, 5 and 6 are determined by X-ray studies.  相似文献   

10.
In addition to well-known dinuclear phenylselenolato palladium complexes, the reaction of [PdCl2(PPh3)2] and NaSePh affords small amounts of novel trinuclear and hexanuclear complexes [Pd3Se(SePh)3(PPh3)3]Cl (1) and [Pd6Cl2Se4(SePh)2(PPh3)6] (2). Complex 1 is triclinic, P1?, a=13.6310(2), b=16.2596(2), c=16.9899(3) Å, α=83.1738(5), β=78.9882(5), γ=78.7635(5)°. Complex 2 is monoclinic, C2/c, a=25.7165(9), b=17.6426(8), c=27.9151(14) Å, β=110.513(2)°. There are no structural forerunners for 1, but the hexanuclear complex 2 is isostructural with [Pd6Cl2Te4(TeR)2(PPh3)6] (R=Ph, C4H3S) that have been observed as one of the products in the oxidative addition of R2Te2 to [Pd(PPh3)4]. Mononuclear palladium complexes may play a significant role as building blocks in the formation of the polynuclear complexes.  相似文献   

11.
Hydrolysis of the 4-alkyliminothiopyrano[2,3-b]pyridinedioles (5) and 4-alkylaminothiopyrano[2,3-b]pyridones (6) resp. with 10% NaOH gives 5,7-dihydroxy-2H-thiopyrano[2,3-b]pyridine-4(3H)-one (7).7 can be obtained in better yield by reaction of 4-dimethylamino-2(1H)-pyridinethione (8) with bistrichlorphenylethylamlonate (2). Aminolysis of7 affords the two isomeric products5 and6. On treatment with hydrazines,7 reacts only to 4-hydrazonoderivatives5. By heating in bromobenzene5d is cyclisized to 1H-5,1,2,6-thiatriaza-acenaphthylen-7-ol (11). On methylation with methyljodide5,6 and7 furnish the 7-methoxyproducts13,14 and12. By heating in 20% NaOH7 is transformed into the 2-thioxo-3-pyridylmethylketone16 A and its tautomer, 2-mercapto-3-pyridylmethylketone16 B. The structures of5,6 and7 are discussed.  相似文献   

12.
2H-Imidazole-4(3H)-thiones (a), available from methyl alkyl and methyl aryl ketones with sulfur and ammonia, react via their corresponding N-sodium compounds or in presence of tert. amines with alkyl and aryl carboxylic acid chlorides to give the corresponding intensely coloured (orange to violett) cryst. 3-acyl-2H-imidazole-4(3H)-thiones4 a-q and6–26. With dicarboxylic acid dichlorides the colourless cryst. N,N′-diacyl-bis-3-imidazoline-5-thiones5 a-d and27–32 are obtained. With carbamic acid chlorides and chloroformic acid esters the corresponding urea (33–35) and urethane derivatives36, 37 are formed. In an analogous way 2H-imidazol-4(3H)-ones react with acid chlorides to 3-acyl-2-imidazol-4(3H)-ones (44–50), which can also be obtained by treating the corresponding 3-acyl-2H-imidazole-4(3H)-thione with KMnO4.  相似文献   

13.
Supramolecular compounds of the compositions {[Cr2(OH)2(H2O)8](C42H42N28O14)2}-(NO3)4·18.75H2O (1) and {[Cr4(OH)6(H2O)12](C48H48N32O16)3(NO3)6·55H2O (2) were synthesized from aqueous solutions of chromium(III) nitrate and the macrocyclic cavitand cucurbit[n]uril (C6n H6n N4n O2n , where n = 7 or 8, respectively). According to the X-ray diffraction study, the polynuclear chromium aqua complexes are disposed in cavities formed by the cucurbit[n]uril molecules and are linked to these molecules through hydrogen bonds between the hydroxo and aqua ligands of the polycations and the portal oxygen atoms of the macrocycles. Compound 1 is the first example of supramolecular compounds of cucurbit[7]uril with metal aqua complexes. The isolation of the supramolecular adduct with cucurbit[8]uril 2 in the single-crystalline state allows the determination of the structure of the tetranuclear chromium aqua complex having an adamantane-like structure, [Cr42-OH)6(H2O)12]6+, which has been previously unknown in the solid state.  相似文献   

14.
Reaction of a triangle Pd(0) complex, Pd3(CNXyl)6 (1; Xyl = 2,6-C6H3Me2), with a dicationic linear trinuclear complex [Pd3(CNXyl)8][PF6]2 (3) afforded a dicationic hexapalladium complex [Pd6(CNXyl)12][PF6]2 (4), while the reaction of 1 with a dicationic dinuclear complex [Pd2(CNXyl)6][PF6]2 (2) resulted in the formation of 3. The molecular structure of the complex 4 was determined by X-ray crystallography and spectroscopic analysis.  相似文献   

15.
The reaction of [Ru(CO)2(PPh3)3] (1) with o-styryldiphenylphophine (SP) (2) gave [Ru(CO)2(PPh3)(SP)] (3) in 83% yield. This styrylphosphine ruthenium complex 3 can also be synthesized by the reaction of [Ru(p-MeOC6H4NN)(CO)2(PPh3)2]BF4 (4) with NaBH4 and 2 in 50% yield. When “Ru(CO)(PPh3)3” generated by the reaction of [RuH2(CO)(PPh3)3] (8) with trimethylvinylsilane reacted with 2, [Ru(CO)(PPh3)2(SP)] (10) was produced in moderate yield as an air sensitive solid. The spectral and X-ray data of these complexes revealed that the coordination geometries around the ruthenium center of both complexes corresponded to a distorted trigonal bipyramid with the olefin occupying the equatorial position and the C-C bonding in the olefin moiety in 3 and 10 contained a significant contribution from a ruthenacyclopropane limiting structure. Complexes 3 and 10 showed catalytic activity for the hydroamination of phenylacetylene 11 with aniline 12. Ruthenium complex 3 in the co-presence of NH4PF6 or H3PW12O40 proves to be a superior catalyst system for this hydroamination reaction. In the case of the reaction using H3PW12O40 as an additive, ketimines (13) was obtained in 99% yield at a ruthenium-catalyst loading of 0.1 mol%. Some aniline derivatives such as 4-methoxy, 4-trifluoromethyl-, and 4-bromoanilines can also be used in this hydroamination reaction.  相似文献   

16.
Heterometallic complexes [RuNO(NO2)4OHCuPy2(H2O)] (I) and [RuNO(NO2)4OHCuPy3] (II) are described structurally for the first time. In complex I, the ruthenium anion is coordinated to the copper atom by the bridging OH group and two bridging nitro groups; in complex II, by the bridging OH group and one bridging nitro group. Dimers are formed in the crystal lattice of complex II due to the interaction of the copper atom and the nitro group of the ruthenium anion in trans position to the bridging NO2 group.  相似文献   

17.
The RhI, RuII, PdI and NiII complexes of the aminobis(phosphonite), PhN(P(OC6H4OMe-o)2)2 (1) are reported. The reactions of 1 with [Rh(COD)Cl]2 in 1:1 and 2:1 molar ratio afford the mono- and diolefin substituted chloro bridged chelate complexes, [(COD)Rh22-Cl)2{PhN(P(OC6H4OMe-o)2)2}] (2) and [Rh(μ2-Cl){PhN(P(OC6H4OMe-o)2)2}]2 (3), respectively. Similarly, the cationic mono- and bis-chelate complexes, [Rh(COD){PhN(P(OC6H4OMe-o)2)2}]OTf (4) and [Rh{PhN(P(OC6H4OMe-o)2)2}2]OTf (5) are obtained by treating 1 with [Rh(COD)Cl]2 in the presence of AgOTf in appropriate ratios. The dinuclear RhI carbonyl complex, [RhCl(CO){μ-PhN(P(OC6H4OMe-o)2)2}]2 (6) is prepared by treating 1 with 0.5 equiv. of [Rh(CO)2Cl]2. Reaction of 1 with cis-[NiBr2(DME)] (DME = 1,2-dimethoxyethane) affords [{PhN(P(OC6H4OMe-o)2)2}NiBr2] (7) whereas with [Ru-(η6-p-cymene)Cl2]2 in refluxing THF medium produces an interesting and rare bimetallic RuII complex, [(η6-p-cymene)Ru(μ2-Cl)3Ru{PhN(P(OC6H4OMe-o)2)2}Cl] (8). Redox condensation of the Pd0 and PdII derivatives with 1 affords the dinuclear PdI complex, [PdBr{μ-PhN(P(OC6H4OMe-o)2)2}]2 (9). The formation and structure of complexes 2-9 are assigned through various spectroscopic and micro analysis data. The molecular structures of 5 and 7-9 are confirmed by single crystal X-ray diffraction studies.  相似文献   

18.
The reaction of the labile compound [Re2(CO)8(CH3CN)2] with 2,3-bis(2-pyridyl)pyrazine in dichloromethane solution at reflux temperature afforded the structural dirhenium isomers [Re2(CO)8(C14H10N4)] (1 and 2), and the complex [Re2(CO)8(C14H10N4)Re2(CO)8] (3). In 1, the ligand is σ,σ′-N,N′-coordinated to a Re(CO)3 fragment through pyridine and pyrazine to form a five-membered chelate ring. A seven-membered ring is obtained for isomer 2 by N-coordination of the 2-pyridyl groups while the pyrazine ring remains uncoordinated. For 2, isomers 2a and 2b are found in a dynamic equilibrium ratio [2a]/[2b]  =  7 in solution, detected by 1H NMR (−50 °C, CD3COCD3), coalescence being observed above room temperature. The ligand in 3 behaves as an 8e-donor bridge bonding two Re(CO)3 fragments through two (σ,σ′-N,N′) interactions. When the reaction was carried out in refluxing tetrahydrofuran, complex [Re2(CO)6(C14H10N4)2] (4) was obtained in addition to compounds 1-3. The dinuclear rhenium derivative 4 contains two units of the organic ligand σ,σ′-N,N′-coordinated in a chelate form to each rhenium core. The X-ray crystal structures for 1 and 3 are reported.  相似文献   

19.
Two novel homobinuclear ytterbium(III) complexes, [Yb2(2AMB)6(H2O)4] · 2C2H6O (I) and Yb2(3AMB)6(H2O)4] · 3H2O (II) (2AMB = 2-aminobenzoic acid, 3AMB = 3-aminobenzoic acid) have been synthesized and characterized by elemental analysis, infrared spectroscopy, thermogravimetric analysis and X-ray crystallography (CIF files CCDC nos. 950103 (I), 921652 (II)). Complex I crystallizes in triclinic space group \(P\bar 1\) and complex II crystallizes in monoclinic space group P21/n. X-ray analysis shows that both complexes (I, II) have the dinuclear structure. The central Yb3+ ions in both complexes are eight-coordinated adopting distorted YbO8 dodecahedral geometry. Each Yb3+ ion is coordinated to two O atoms from bridging carboxylate, four O atoms from the chelating carboxylate ligands and two O atoms of water molecules. The crystal structure of I and II are stabilized by N-H…O, O-H…O, O-H…N, and C-H…O hydrogen bonds, C-H…π interactions and weak π-π stacking interactions.  相似文献   

20.
The first example of isonicotinic acid compounds with infinite mercury halide chains, [HgCl2(C6NO2H5)]n n[HgCl2]n(C6NO2H5) (1), was synthesized through hydrothermal reactions and structurally characterized by X-ray single crystal diffraction. Compound 1 features a one-dimensional (1-D) motif, based on infinite 1-D [HgCl2(C6NO2H5)]n chains, neutral HgCl2 moieties and isolated isonicotinic acid molecules. The [HgCl2(C6NO2H5)]n chains, HgCl2 moieties and isonicotinic acid molecules are interlinked by hydrogen bonds and π-π interactions to give a two-dimensional supramolecular layer. Photoluminescent investigation reveals that the title compound exhibits a strong emission in blue region. The emission band is identified as the π -π* transitions of the isonicotinic acid moieties.  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