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.
A solution approach based on Au(CN)2? chemistry is reported for the formation of nanoparticles. The covalent character of the Au(CN)2? precursor was exploited in the formation of sub‐10 nm nanospheres (≈2.4 nm) and highly monodisperse icosahedral Au nanoparticles (≈8 nm) at room temperature in a one‐pot aqueous synthesis. The respective spherical and icosahedral Au morphologies can be controlled by either the absence or presence of the polymer polyvinylpyrrolidone (PVP). Using Au(CN)2? as a metal ion source, our findings suggest that the addition of citrate ions is necessary to enhance the particle formation rate as well as to generate a more homogeneous colloidal dispersion. Because of the presence of oxygen and the operation of a CN? etching process associated with Au(CN)2? complex formation, an interesting reversible formation–dissolution process was observed, which allowed us to repeatedly prepare spherical and icosahedral Au nanoparticles. Time‐dependent TEM images and UV/Vis spectra were carefully acquired to study the reversibility of this formation–dissolution process. In view of the accompanying generation of toxic cyanide anions, we have developed a protocol to recycle cyanide in the presence of citrate ions through ferric ferrocyanide formation. After completion of particle formation, the residual solutions containing citrate ions and cyanide ions were processed to stain iron oxide nanoparticles endocytosized in cells. Additionally, the as‐prepared 8 nm Au icosahedra could be isolated and grown to larger 57 nm‐sized icosahedra using the seed‐mediated growth approach.  相似文献   

2.
Organometallic multi‐decker sandwich complexes containing f‐elements remain rare, despite their attractive magnetic and electronic properties. The reduction of the CeIII siloxide complex, [KCeL4] ( 1 ; L=OSi(OtBu)3), with excess potassium in a THF/toluene mixture afforded a quadruple‐decker arene‐bridged complex, [K(2.2.2‐crypt)]2[{(KL3Ce)(μ‐η66‐C7H8)}2Ce] ( 3 ). The structure of 3 features a [Ce(C7H8)2] sandwich capped by [KL3Ce] moieties with a linear arrangement of the Ce ions. Structural parameters, UV/Vis/NIR data, and DFT studies indicate the presence of CeII ions involved in δ bonding between the Ce cations and toluene dianions. Complex 3 is a rare lanthanide multi‐decker complex and the first containing non‐classical divalent lanthanide ions. Moreover, oxidation of 1 by AgOTf (OTf=O3SCF3) yielded the CeIV complex, [CeL4] ( 2 ), showing that siloxide ligands can stabilize Ce in three oxidation states.  相似文献   

3.
The title compound, diaqua‐1κO,3κO‐di‐μ‐hydroxido‐1:2κ2O:O,2:3κ2O:O‐di‐μ‐methacrylato‐1:2κ2O:O′,2:3κ2O:O′‐bis(1,10‐phenanthroline)‐1κ2N,N′;3κ2N,N′‐tricopper(II) dinitrate dihydrate, [Cu3(C4H5O2)2(OH)2(C12H8N2)2(H2O)2](NO3)2·2H2O, has the central Cu atom on an inversion centre. The three CuII atoms are in a linear arrangement linked by methacrylate and hydroxide groups. The coordination environments of the CuII ions are five‐coordinated distorted square‐pyramidal for the outer Cu atoms and four‐coordinated square‐planar for the central Cu atom. All nitrate ions, hydroxide groups and water molecules are linked by hydrogen bonds, forming a linear structure. The complex exhibits ferromagnetic exchange coupling, which is helpful in elucidating magnetic interactions between copper ions and other metallic ions in heteronuclear complexes.  相似文献   

4.
A new bimetallic FeII–CuII complex was synthesized, characterized, and applied as a selective and sensitive sensor for cyanide detection in water. This complex is the first multifunctional device that can simultaneously detect cyanide ions in real water samples, amplify the colorimetric signal upon detection for naked‐eye recognition at the parts‐per‐million (ppb) level, and convert the toxic cyanide ion into the much safer cyanate ion in situ. The mechanism of the bimetallic complex for high‐selectivity recognition and signaling toward cyanide ions was investigated through a series of binding kinetics of the complex with different analytes, including CN?, SO42?, HCO3?, HPO42?, N3?, CH3COO?, NCS?, NO3?, and Cl? ions. In addition, the use of the indicator/catalyst displacement assay (ICDA) is demonstrated in the present system in which one metal center acts as a receptor and inhibitor and is bridged to another metal center that is responsible for signal transduction and catalysis, thus showing a versatile approach to the design of new multifunctional devices.  相似文献   

5.
Redox‐inactive metal ions are one of the most important co‐factors involved in dioxygen activation and formation reactions by metalloenzymes. In this study, we have shown that the logarithm of the rate constants of electron‐transfer and C−H bond activation reactions by nonheme iron(III)–peroxo complexes binding redox‐inactive metal ions, [(TMC)FeIII(O2)]+‐Mn + (Mn +=Sc3+, Y3+, Lu3+, and La3+), increases linearly with the increase of the Lewis acidity of the redox‐inactive metal ions (ΔE ), which is determined from the gzz values of EPR spectra of O2.−‐Mn + complexes. In contrast, the logarithm of the rate constants of the [(TMC)FeIII(O2)]+‐Mn + complexes in nucleophilic reactions with aldehydes decreases linearly as the ΔE value increases. Thus, the Lewis acidity of the redox‐inactive metal ions bound to the mononuclear nonheme iron(III)–peroxo complex modulates the reactivity of the [(TMC)FeIII(O2)]+‐Mn + complexes in electron‐transfer, electrophilic, and nucleophilic reactions.  相似文献   

6.
The structures of five metal complexes containing the 4‐oxo‐4H‐pyran‐2,6‐dicarboxylate dianion illustrate the remarkable coordinating versatility of this ligand and the great structural diversity of its complexes. In tetraaquaberyllium 4‐oxo‐4H‐pyran‐2,6‐dicarboxylate, [Be(H2O)4](C7H2O6), (I), the ions are linked by eight independent O—H...O hydrogen bonds to form a three‐dimensional hydrogen‐bonded framework structure. Each of the ions in hydrazinium(2+) diaqua(4‐oxo‐4H‐pyran‐2,6‐dicarboxylato)calcate, (N2H6)[Ca(C7H2O6)2(H2O)2], (II), lies on a twofold rotation axis in the space group P2/c; the anions form hydrogen‐bonded sheets which are linked into a three‐dimensional framework by the cations. In bis(μ‐4‐oxo‐4H‐pyran‐2,6‐dicarboxylato)bis[tetraaquamanganese(II)] tetrahydrate, [Mn2(C7H2O6)2(H2O)8]·4H2O, (III), the metal ions and the organic ligands form a cyclic centrosymmetric Mn2(C7H2O6)2 unit, and these units are linked into a complex three‐dimensional framework structure containing 12 independent O—H...O hydrogen bonds. There are two independent CuII ions in tetraaqua(4‐oxo‐4H‐pyran‐2,6‐dicarboxylato)copper(II), [Cu(C7H2O6)(H2O)4], (IV), and both lie on centres of inversion in the space group P; the metal ions and the organic ligands form a one‐dimensional coordination polymer, and the polymer chains are linked into a three‐dimensional framework containing eight independent O—H...O hydrogen bonds. Diaqua(4‐oxo‐4H‐pyran‐2,6‐dicarboxylato)cadmium monohydrate, [Cd(C7H2O6)(H2O)2]·H2O, (V), forms a three‐dimensional coordination polymer in which the organic ligand is coordinated to four different Cd sites, and this polymer is interwoven with a complex three‐dimensional framework built from O—H...O hydrogen bonds.  相似文献   

7.
2‐Mercaptopyridine N ‐oxide (pyrithione, PTOH) along with several transition metal ions forms coordination compounds displaying notable biological activities. Gas‐phase complexes formed between pyrithione and manganese (II), cobalt (II), nickel (II), copper (II), and zinc (II) were investigated by infusion in the electrospray source of a quadrupole‐time of flight mass spectrometer. Remarkably, positive ion mode spectra displayed the singly charged metal adduct ion [C10H8MN2O2S2]2+ ([M(PTO)2]+• or [M(DPTO)]+•), where DPTO is dipyrithione, 2,2′‐dithiobis(pyridine N ‐oxide), among the most abundant peaks, implying a change in the oxidation state of whether the metal ion or the ligands. In addition, doubly charged ions were recognized as metal adduct ions containing DPTO ligands, [M(DPTO)n]2+. Generation of [M(PTO)2]+• / [M(DPTO)]+• could be traced by CID of [M(DPTO)2]2+, by observation of the sequential losses of a charged (PTO+) and a radical (PTO) deprotonated pyrithione ligand. The fragmentation pathways of [M(PTO)2]+• / [M(DPTO)]+• were compared among the different metal ions, and some common features were noticed. Density functional theory (DFT) calculations were employed to study the structures of the observed adduct ions, and especially, to decide in the adduct ion [M(PTO)2]+• / [M(DPTO)]+• whether the ligands are 2 deprotonated pyrithiones or a single dipyrithione as well as the oxidation state of the metal ion in the complex. Characterization of gas‐phase pyrithione metal ion complexes becomes important, especially taking into account the presence of a redox‐active ligand in the complexes, because redox state changes that produce new species can have a marked effect on the overall toxicological/biological response elicited by the metal system.  相似文献   

8.
Complexes of 4,10‐bis(phosphonomethyl)‐1,4,7,10‐tetraazacyclododecane‐1,7‐diacetic acid (trans‐H6do2a2p, H6 L ) with transition metal and lanthanide(III) ions were investigated. The stability constant values of the divalent and trivalent metal‐ion complexes are between the corresponding values of H4dota and H8dotp complexes, as a consequence of the ligand basicity. The solid‐state structures of the ligand and of nine lanthanide(III) complexes were determined by X‐ray diffraction. All the complexes are present as twisted‐square‐antiprismatic isomers and their structures can be divided into two series. The first one involves nona‐coordinated complexes of the large lanthanide(III) ions (Ce, Nd, Sm) with a coordinated water molecule. In the series of Sm, Eu, Tb, Dy, Er, Yb, the complexes are octa‐coordinated only by the ligand donor atoms and their coordination cages are more irregular. The formation kinetics and the acid‐assisted dissociation of several LnIII–H6 L complexes were investigated at different temperatures and compared with analogous data for complexes of other dota‐like ligands. The [Ce( L )(H2O)]3? complex is the most kinetically inert among complexes of the investigated lanthanide(III) ions (Ce, Eu, Gd, Yb). Among mixed phosphonate–acetate dota analogues, kinetic inertness of the cerium(III) complexes is increased with a higher number of phosphonate arms in the ligand, whereas the opposite is true for europium(III) complexes. According to the 1H NMR spectroscopic pseudo‐contact shifts for the Ce–Eu and Tb–Yb series, the solution structures of the complexes reflect the structures of the [Ce(H L )(H2O)]2? and [Yb(H L )]2? anions, respectively, found in the solid state. However, these solution NMR spectroscopic studies showed that there is no unambiguous relation between 31P/1H lanthanide‐induced shift (LIS) values and coordination of water in the complexes; the values rather express a relative position of the central ions between the N4 and O4 planes.  相似文献   

9.
The kinetics and thermodynamics of O2 addition to CoII complexes containing the simple triamine ligand (L) diethylenetriamine (=N‐(2‐aminoethyl)ethane‐1,2‐diamine; dien) or N,N″‐dimethyldiethylenetriamine (=N‐methyl‐N′‐[2‐(methylamino)ethyl]ethane‐1,2‐diamine; dmdien) in the aprotic solvent dimethyl sulfoxide (DMSO) were studied by UV/VIS spectrophotometry, potentiometry, and O2 absorption measurements. A parallel investigation on the anaerobic formation of CoII complexes with dmdien, as well as on their reactivity towards O2, was carried out in aqueous 0.1M NaClO4 solution. [CoL]2+ and [CoL2]2+ were the common species formed under anaerobic conditions in both aqueous and DMSO solutions. Under aerobic conditions, O2 adducts of different stoichiometry were formed: a superoxo complex [CoL2O2]2+ in DMSO and dimeric species in H2O. The role of the reaction medium as well as effects of N‐alkylation of the triamine ligand in the formation and reactivity of the [CoII(triamine)] complexes are discussed.  相似文献   

10.
A novel 3D polymeric heteropolynuclear sodium(I) lead(II) complex containing different ligands, [NaPb(ClO4)(en)(NO2)2] was synthesized and characterized by elemental analysis and IR, and 1H‐, 13C‐, and 207Pb‐NMR spectroscopy. The single‐crystal X‐ray data of [NaPb(ClO4)(en)(NO2)2]n established that the complex is a three‐dimensional polymer, [(en)Pb(μ3‐ONO)2Na(μ3‐ONO)2Na(μ‐O2ClO2)Na]n. The Pb and Na atoms have four‐ and eight‐coordinate geometry, respectively. The lone pair of electrons at the PbII atom is ‘stereochemically active’.  相似文献   

11.
The spectroscopic behavior of ionic Eu3+ or Tb3+ complexes of an aromatic carboxyl‐functionalized organic salt as well as those of the hybrid materials derived from adsorption of the ionic complexes on Laponite clay are reported. X‐ray diffraction (XRD) patterns suggest that the complexes are mainly adsorbed on the outer surfaces of the Laponite disks rather than intercalated within the interlayer spaces. Photophysical data showed that the energy‐transfer efficiency from the ligand to Eu3+ ions in the hybrid material is increased remarkably with respect to the corresponding ionic complex. The hybrid material containing the Eu3+ complex shows bright red emission from the prominent 5D07F2 transition of Eu3+ ions, and that containing the Tb3+ complex exhibits bright green emission due to the dominant 5D47F5 transition of Tb3+ ions.  相似文献   

12.
Electrospray ionization quadrupole ion trap mass spectrometry (ESI-QIT/MS) of the ammonium cerium(III) phosphomolybdate complex (NH4)11[Ce(III)(PMo11O39)2] in aqueous media has revealed a concentration-dependent behavior. Under fixed instrumental parameters, the Ce-containing polyoxomolybdate complexes H2Ce(III)P2Mo22O753? and Ce(III)PMo11O382? are the primary species present at 11 mM (pH = 4.3); at 0.7 mM (pH = 3.6), Ce(III)PMo10O352? is the predominant species, Ce(III)PMo11O382? is quite diminished, and H2Ce(III)P2Mo22O753? is absent. As a result of the complex isotopic fingerprints from multiple molybdenums, compositions of such ions are difficult to assign—successive collision induced dissociation (CID) of large ions produced smaller ions for which calculated and experimental isotopic patterns could be compared. The oxidation state of Ce and the number of counter cations on negative complexes was discerned from spectra of ions containing 1H+ and 7Li+. The overall result is an ESI method applicable to phosphomolybdate complexes containing redox sensitive f-block metal ions such as Ce(IV) and Pu(III/IV). Dissociation studies also gave insight into favored fragmentation pathways, and generated gas ions with empirical formulae similar to known condensed-phase ions. Deconvolution of concentration- and pH-dependent solution behavior via ESI/MS and 31P NMR spectroscopy showed speciation dependent on solution concentration, not on pH.  相似文献   

13.
Formation of 8‐oxoguanine (8OG) from guanine in biological systems is known to cause lethal mutation and cancer. It has been suggested earlier, on the basis of experimental studies, that the oxygen molecule in its lowest singlet excited state (1O2) plays an important role in the formation of 8OG. In order to understand the possible mechanisms in this context, B3LYP/6‐31+G* and MP2/6‐31+G* calculations were carried out on the structures and stabilities of different molecules and complexes involved in the formation of 8OG. All the molecules, complexes, and transition states studied in the present report were solvated in aqueous media. Guanine has been found to make a strong complex with 1O2 with the latter species located above the imidazole ring plane, and the complex of guanine with 3O2 is much weaker than that with 1O2. Transition state calculations were carried out to study formation of 7,8‐dihydro,8‐hydroxyguanine (8OHG) and 2‐oxo‐imidazole. It has been shown that 8OG can be formed in two different ways: (i) due to interaction of the radical cation of guanine with O where 8OHG complexed with 1O2 would occur as an intermediate, and (ii) due to interaction of guanine with 1O2 leading to the formation of guanine hydroperoxide that would react with a water molecule in the presence of two 1O2 molecules serving as a source of energy to overcome the barrier. It is shown that because the interaction strengths of 3O2 and 1O2 with other molecules, e.g., guanine, are very different, a crossing of their potential energy surfaces takes place in both gas phase and aqueous media, as a result of which the lifetime of 1O2 is strongly decreased. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

14.
Two distinct polymorphs of bis(μ2‐methylquinolin‐8‐olato)‐κ3N,O:O3O:N,O‐bis[(isothiocyanato‐κN)lead(II)], [Pb2(C10H8NO)2(NCS)2], (I), forming dinuclear complexes from a methanolic solution containing lead(II) nitrate, 2‐methylquinolin‐8‐ol (M‐Hq) and KSCN, crystallized concomitantly as colourless prisms [form (Ia)] and long thin colourless needles [form (Ib)]. In both cases, the complexes lie across a centre of inversion. The polymorphs differ substantially in their conformation and in their interactions, viz. Pb...S and π–π for form (Ia) and Pb...S, Pb...π and C—H...π for form (Ib).  相似文献   

15.
AgI‐containing coordination complexes have attracted attention because of their photoluminescence properties and antimicrobial activities and, in principle, these properties depend on the nature of the structural topologies. A novel two‐dimensional silver(I) complex with the anti‐inflammatory diclofenac molecule, namely bis{μ‐2‐[2‐(2,6‐dichloroanilino)phenyl]acetato‐κ3O,O′:O}bis(μ‐2,5‐dimethylpyrazine‐κ2N:N′)silver(I), [Ag2(C14H10Cl2NO2)2(C6H8N2)]n, (I), has been synthesized and characterized by single‐crystal X‐ray diffraction, revealing that the AgI ions are chelated by the carboxylate groups of the anionic 2‐[2‐(2,6‐dichloroanilino)phenyl]acetate (dicl) ligand in a μ3‐η12 coordination mode. Each dicl ligand links three AgI atoms to generate a one‐dimensional infinite chain. Adjacent chains are connected through 2,5‐dimethylpyrazine (dmpyz) ligands to form a two‐dimensional layer structure parallel to the crystallographic bc plane. The layers are further connected by C—H…π interactions to generate a three‐dimensional supramolecular structure. Additionally, the most striking feature is that the structure contains an intramolecular C—H …Ag anagostic interaction. Furthermore, the title complex has been tested for its in vitro antibacterial activity and is determined to be highly effective on the studied microorganisms.  相似文献   

16.
The geometries and energetics of complexes of Hg(II) and Pb(II) with sulfur‐ and aminopyridine‐containing chelating resin including crosslinked polystyrene immobilizing 2‐aminopyridine via sulfur‐containing (PVBS‐AP), sulfoxide‐containing (PVBSO‐AP), and sulfone‐containing (PVBSO2‐AP) spacer arms have been investigated theoretically, and thus interactions of the metal ions with chelating resins were evaluated. The results indicate that PVBS‐AP behaves as a tridentate ligand to coordinate with the metal ions by S and two N atoms to form chelating compounds with S atom playing a dominant role in the coordination, whereas PVBSO‐AP and PVBSO2‐AP interact with metal cations, respectively, in a tricoordinate manner by O and two N atoms forming chelating complexes. Furthermore, it is revealed that O and N2 atoms of PVBSO‐AP are the main contributor of coordination to Hg(II), whereas N2 atom of PVBSO2‐AP is mainly responsible for the coordination to Hg(II). For PVBSO‐AP‐Pb2+ and PVBSO2‐AP‐Pb2+ complex, the coordination is dominated by the synergetic effect of N1, N2, and O atoms. Natural bond orbital and second‐order perturbation analyses suggest that the charge transfer from the chelating resins to metal ions is mainly dominated by the interactions of lone pair of electrons of the donor atoms with the unoccupied orbitals of metal ions. Hg(II) complexes exhibit larger binding energies than the corresponding Pb(II) complexes, implying the chelating resins exhibit higher affinity toward Hg(II), which is consistent with the experimental results. Combined the theoretical and experimental results, further understanding of the structural information of the complexes and the coordination mechanism was achieved. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

17.
1‐Benzofuran‐2,3‐dicarboxylic acid (C10H6O5) is a dicarboxylic acid ligand which can readily engage in organometallic complexes with various metal ions. This ligand is characterized by an intramolecular hydrogen bond between the two carboxyl residues, and, as a monoanionic species, readily forms supramolecular adducts with different organic and inorganic cations. These are a 1:1 adduct with the dimethylammonium cation, namely dimethylammonium 3‐carboxy‐1‐benzofuran‐2‐carboxylate, C2H8N+·C10H5O5, (I), a 2:1 complex with Cu2+ ions in which four neutral imidazole molecules also coordinate the metal atom, namely bis(3‐carboxy‐1‐benzofuran‐2‐carboxylato‐κO3)tetrakis(1H‐imidazole‐κN3)copper(II), [Cu(C10H5O5)2(C3H4N2)4], (II), and a 4:1 adduct with [La(H2O)7]3+ ions, namely heptaaquabis(3‐carboxy‐1‐benzofuran‐2‐carboxylato‐κO3)lanthanum 3‐carboxy‐1‐benzofuran‐2‐carboxylate 1‐benzofuran‐2,3‐dicarboxylic acid solvate tetrahydrate, [La(C10H5O5)2(H2O)7](C10H5O5)·C10H6O5·4H2O, (III). In the crystal structure, complex (II) resides on inversion centres, while complex (III) resides on axes of twofold rotation. The crystal packing in all three structures reveals π–π stacking interactions between the planar aromatic benzofuran residues, as well as hydrogen bonding between the components. The significance of this study lies in the first crystallographic characterization of the title framework, which consistently exhibits the presence of an intramolecular hydrogen bond and a consequent monoanionic‐only nature. It shows further that the anion can coordinate readily to metal cations as a ligand, as well as acting as a monovalent counter‐ion. Finally, the aromaticity of the flat benzofuran residue provides an additional supramolecular synthon that directs and facilitates the crystal packing of compounds (I)–(III).  相似文献   

18.
Two complexes containing (nBu)2Sn4O4L4 (L = salicylaminoaryl alcohols) were synthesized and characterized by elemental analysis and 1H, 13C, and 119Sn NMR as well as IR spectroscopy. The crystal structure of complex 2 . 2 was determined by X‐ray crystallography. It consists of three, four‐member Sn2O2 rings in a ladder‐like, structural arrangement with four tin centers, bridged by four, three‐coordinate μ3‐oxygen atoms. The dimeric distannoxane have two different, pentacoordinate tin atoms, where the Schiff base acts as monodentate ligand. The complexes were subjected to a thiobarbituric acid, reactive substance (TBARS) assay, which displayed a higher antioxidant activity than the α‐tocopherol and Butylated hydroxy toluene (BHT), used as positive controls.  相似文献   

19.
Square‐planar complexes are commonly formed by transition metal ions having a d8 electron configuration. Planar cyanometallate anions have been used extensively as design elements in supramolecular coordination systems. In particular, square‐planar tetracyanometallate(II) ions, i.e. [M(CN)4]2− (MII = Ni, Pd or Pt), are used as good building blocks for bimetallic Hofmann‐type assemblies and their analogues. Square‐planar tetracyanonickellate(II) complexes have been extensively developed with N‐donor groups as additional co‐ligands, but studies of these systems using O‐donor ligands are scarce. A new cyanide‐bridged CuII–NiII heterometallic compound, poly[[diaquatetra‐μ2‐cyanido‐κ8C:N‐nickel(II)copper(II)] monohydrate], {[CuIINiII(CN)4(H2O)2]·H2O}n, has been synthesized and characterized by X‐ray single‐crystal diffraction analyses, vibrational spectroscopy (FT–IR), thermal analysis, electron paramagnetic resonance (EPR) and magnetic moment measurements. The structural analysis revealed that it has a two‐dimensional grid‐like structure built up of cationic [Cu(H2O)2]2+ and anionic [Ni(CN)4]2− units connected through bridging cyanide ligands. The overall three‐dimensional supramolecular network is expanded by a combination of interlayer O—H…N and intralayer O—H…O hydrogen‐bond interactions. The first decomposition reactions take place at 335 K under a static air atmosphere, which illustrates the existence of guest water molecules in the interlayer spaces. The electron paramagnetic resonance (EPR) spectrum confirms that the CuII cation has an axial coordination symmetry and that the unpaired electrons occupy the d orbital. In addition, magnetic investigations showed that antiferromagnetic interactions exist in the CuII atoms through the diamagnetic [Ni(CN)4]2− ion.  相似文献   

20.
Two new arene inverted‐sandwich complexes of uranium supported by siloxide ancillary ligands [K{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 3 ) and [K2{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 4 ) were synthesized by the reduction of the parent arene‐bridged complex [{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 2 ) with stoichiometric amounts of KC8 yielding a rare family of inverted‐sandwich complexes in three states of charge. The structural data and computational studies of the electronic structure are in agreement with the presence of high‐valent uranium centers bridged by a reduced tetra‐anionic toluene with the best formulation being UV–(arene4?)–UV, KUIV–(arene4?)–UV, and K2UIV–(arene4?)–UIV for complexes 2 , 3 , and 4 respectively. The potassium cations in complexes 3 and 4 are coordinated to the siloxide ligands both in the solid state and in solution. The addition of KOTf (OTf=triflate) to the neutral compound 2 promotes its disproportionation to yield complexes 3 and 4 (depending on the stoichiometry) and the UIV mononuclear complex [U(OSi(OtBu)3)3(OTf)(thf)2] ( 5 ). This unprecedented reactivity demonstrates the key role of potassium for the stability of these complexes.  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