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 31 毫秒
1.
Cis-/Trans-Isomerism of Bis-(trisalkoxy)-hexavanadates: cis-Na2[V O7(OH)6{(OCH2)3CCH2OH}2] · 8 H2O, cis-(CN3H6)3[VIVV O13{(OCH2)3CCH2OH}2] · 4.5 H2O and trans-(CN3H6)2[V O13{(OCH2)3CCH2OH}2] · H2O Polyoxovanadates with distorted Lindquist-structure, in which six of the twelve μ2-oxygen atoms are formally replaced by the oxygen atoms of two coordinated pentaerythritol ligands, can be prepared by a simple method in an aqueous medium. The “fully reduced”, six-fold protonated compound cis-Na2[VO7(OH)6{(OCH2)3CCH2OH}2] · 8 H2O ( 1 ), the mixed valence species cis-(CN3H6)3[VIVVO13{(OCH2)3CCH2OH}2] · 4.5 H2O ( 2 ) containing one localized VIV centre and the “fully oxidized” compound trans-(CN3H6)2[VO13{(OCH2)3CCH2 · OH}2] · H2O ( 3 ) have been synthesized and characterized by UV/VIS-, IR- and EPR-spectroscopy, by magnetic measurements, cyclic voltammetry and by a single-crystal X-ray structure analysis. The organic {(CH2)3CCH2OH}3+-groups tend to cap the triangular faces formed by μ2-oxygen atoms of the central approximately octahedral {V6O19}-unit. Therefore the anions of bis-(trisalkoxy)-hexavanadates can exist in a trans-form as well as in an isomeric cis-form referring to a “basic” plane of four vanadium atoms of the {V6}-octahedron. The different relative positions of the ligands have a significant influence on the redox potentials of the compounds. For structural details see “Inhaltsübersicht”.  相似文献   

2.
Four aromatic hybrid Anderson polyoxomolybdates with Fe3+ or Mn3+ as the central heteroatom have been synthesized by using a pre‐functionalization protocol and characterized by using single‐crystal X‐ray diffraction, FTIR, ESI‐MS, 1H NMR spectroscopy, and elemental analysis. Structural analysis revealed the formation of (TBA)3[FeMo6O18{(OCH2)3CNHCOC6H5}2] ? 3.5 ACN ( TBA‐FeMo6‐bzn ; TBA=tetrabutylammonium, ACN=acetonitrile, bzn=TRIS‐benzoic acid alkanolamide, TRIS?R=(HOCH2)3C?R)), (TBA)3[FeMo6O18{(OCH2)3CNHCOC8H7}2] ? 2.5 ACN ( TBA‐FeMo6‐cin ; cin=TRIS‐cinnamic acid alkanolamide), (TBA)3[MnMo6O18{(OCH2)3CNHCOC6H5}2] ? 3.5 ACN ( TBA‐MnMo6‐bzn ), and (TBA)3[MnMo6O18{(OCH2)3CNHCOC8H7}2] ? 2.5 ACN ( TBA‐MnMo6‐cin ). To make these four compounds applicable in biological systems, an ion exchange was performed that gave the water‐soluble (up to 80 mM ) sodium salts Na3[FeMo6O18{(OCH2)3CNHCOC6H5}2] ( Na‐FeMo6‐bzn ), Na3[FeMo6O18{(OCH2)3CNHCOC8H7}2] ( Na‐FeMo6‐cin ), Na3[MnMo6O18{(OCH2)3CNHCOC6H5}2] ( Na‐MnMo6‐bzn ), and Na3[MnMo6O18{(OCH2)3CNHCOC8H7}2] ( Na‐MnMo6‐cin ). The hydrolytic stability of the sodium salts was examined by applying ESI‐MS in the pH range of 4 to 9. Sodium dodecylsulfate–polyacrylamide gel electrophoresis (SDS‐PAGE) showed that human and bovine serum albumin (HSA and BSA) remain intact in solutions that contain up to 100 equivalents of the sodium salts over more than 4 d at 20 °C. Tryptophan (Trp) fluorescence quenching was applied to study the interactions between the sodium salts and HSA and BSA at pH 5.5 and 7.4. The quenching constants were extracted by using Stern–Volmer analysis, which suggested the formation of a 1:1 POM–protein complex in all samples. It is suggested that the aromatic hybrid POM approaches subdomain IIA of HSA and exhibits hydrophobic interactions with its hydrophobic tails, whereas the Anderson core is stabilized through electrostatic interactions with polar amino acid side chains from, for example, subdomain IB.  相似文献   

3.
Tantalum complexes [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(CH2NMe2)?CH)py}] ( 4 ) and [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(CH2NH2)?CH)py}] ( 5 ), which contain modified alkoxide pincer ligands, were synthesized from the reactions of [TaCp*Me{κ3N,O,O‐(OCH2)(OCH)py}] (Cp*=η5‐C5Me5) with HC?CCH2NMe2 and HC?CCH2NH2, respectively. The reactions of [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(Ph)?CH)py}] ( 2 ) and [TaCp*Me{κ4C,N,O,O‐(OCH2)(OCHC(SiMe3)?CH)py}] ( 3 ) with triflic acid (1:2 molar ratio) rendered the corresponding bis‐triflate derivatives [TaCp*(OTf)23N,O,O‐(OCH2)(OCHC(Ph)?CH2)py}] ( 6 ) and [TaCp*(OTf)23N,O,O‐(OCH2)(OCHC(SiMe3)?CH2)py}] ( 7 ), respectively. Complex 4 reacted with triflic acid in a 1:2 molar ratio to selectively yield the water‐soluble cationic complex [TaCp*(OTf){κ4C,N,O,O‐(OCH2)(OCHC(CH2NHMe2)?CH)py}]OTf ( 8 ). Compound 8 reacted with water to afford the hydrolyzed complex [TaCp*(OH)(H2O){κ3N,O,O‐(OCH2)(OCHC(CH2NHMe2)?CH2)py}](OTf)2 ( 9 ). Protonation of compound 8 with triflic acid gave the new tantalum compound [TaCp*(OTf){κ4C,N,O,O‐(OCH2)(HOCHC(CH2NHMe2)?CH)py}](OTf)2 ( 10 ), which afforded the corresponding protonolysis derivative [TaCp*(OTf)23N,O,O‐(OCH2)(HOCHC(CH2NHMe2)?CH2)py}](OTf) ( 11 ) in solution. Complex 8 reacted with CNtBu and potassium 2‐isocyanoacetate to give the corresponding iminoacyl derivatives 12 and 13 , respectively. The molecular structures of complexes 5 , 7 , and 10 were established by single‐crystal X‐ray diffraction studies.  相似文献   

4.
Molybdenum(II) Halide Clusters with six Alcoholate Ligands: (C18H36N2O6Na)2[Mo6Cl8(OCH3)6] · 6CH3OH and (C18H36N2O6Na)2[Mo6Cl8(OC6H5)6] . The reaction of Na2[Mo6Cl8(OCH3)6] and 2,2,2-crypt yields (C18H36N2O6Na)2[Mo6Cl8(OCH3)6] · 6 CH3OH ( 1 ), which is converted to (C18H36N2O6Na)2[Mo6Cl8(OC6H5)6] ( 2 ) by metathesis with phenol. According to single crystal structure determinations ( 1 : P3 1c, a=14.613(3) Å, c=21.036(8) Å; 2 : P3 1c, a=15.624(1) Å, c=19.671(2) Å) the compounds contain anionic clusters [Mo6Cl8i(ORa)6]2? ( 1 : d(Mo—Mo) 2.608(1) Å to 2.611(1) Å, d(Mo—Cl) 2.489(1) Å to 2.503(1) Å, d(Mo—O) 2.046(4) Å; 2 : d(Mo—Mo) 2.602(3) Å to 2.608(3) Å, d(Mo—Cl) 2.471(5) Å to 2.4992(5) Å, d(Mo—O) 2.091(14) Å). Electronic interactions of the halide cluster and the phenolate ligands in [Mo6Cl8(OC6H5)6]2? is investigated by means of UV/VIS spectroscopy and EHMO calculations.  相似文献   

5.
Monocyclometalated compound [Rh2{(C8H4S)P(C8H5S)2}(CH3CO2H)2(O2CCH3)3] ( 1 a ) and bis‐cyclometalated compound [Rh2{(C8H4S)P(C8H5S)2}2(CH3CO2H)2(O2CCH3)2] ( 2 a ) have been isolated from the reaction of dirhodium tetraacetate and tris(2‐benzo[b]thienyl)phosphine ( 2 BTP ) using low acidic solutions. By contrast, in pure acetic acid the reaction of Rh2(O2CCH3)4 with 2 BTP and tris(2‐thienyl)phosphine ( 2 TP ), followed by replacement of the axial acetate ligands by chlorides, led to [Rh2{(2‐C8H5 S )P(2‐C8H5S)2}2Cl2(O2CCH3)2] ( 3 b ) and [Rh2{(2‐C4H3 S )P(C4H3S)2}2Cl2(O2CCH3)2] ( 5 b ), respectively. These new dirhodium(II) compounds possess equatorial bridging ligands in a phosphorous–sulfur (P,S) coordination mode. The reversible switching between the P,C and P,S bonding mode of the phosphine has been studied in the monocyclometalated [Rh2{(C4H2S)P(C4H3S)2}(CH3CO2H)2(O2CCH3)3] ( 6 a ), which was selectively transformed into compound [Rh2{(2‐C4H3 S )P(C4H3S)2}(CF3SO3)(CH3CO2H)(O2CCH3)3] ( 7 c ) in triflic acid media. Remarkably, compound 7 c reverts to the starting compound 6 a upon treatment with sodium acetate. Theoretical DFT calculations for both the P,C/P,S rearrangement and the base‐promoted reversion have been performed to explain the experimental findings. Data suggest the P,C/P,S rearrangement occurs by means of a “concerted protonation–demetalation mechanism” followed by η2 coordination of the thienyl ring and subsequent isomerization to the S‐η1‐coordination mode. In the reversion reaction, the base coordinated at the axial position would promote a concerted metalation–deprotonation mechanism.  相似文献   

6.
The new transition metal oxo‐thiostannate {[Ni(cyclen)]6[Sn6S12O2(OH)6]} · 2(ClO4) · 19H2O ( 1 ) was prepared under hydrothermal conditions using Na4SnS4 · 14H2O and [Ni(cyclen)](ClO4)2 as reactants. In the crystal structure the rare [Sn6S12O2(OH)6]10– anion is observed, which is composed of SnS2O(OH)3 and SnS4O2 octahedra, and SnS4 tetrahedra sharing edges and corners. The anion is expanded by six Ni2+ centered complexes via Ni–S and Ni–OH bonds. The photocatalytic properties for the visible light driven hydrogen evolution reaction shows that 26.6 mmol · g–1 H2 were evolved after 3 h.  相似文献   

7.
Three unprecedented 2D architectures made up of sandwich‐type tetra‐metal‐substituted polyoxotungstates and transition‐metal complexes, [Cu(dien)(H2O)]2{[Cu(dien)(H2O)]2‐[Cu(dien)(H2O)2]2[Cu4(SiW9O34)2]}? 5H2O ( 1 ; dien=diethylenetriamine), [Zn(enMe)2(H2O)]2{[Zn(enMe)2]2[Zn4‐ (HenMe)2(PW9O34)2]}?8H2O ( 2 ; enMe =1,2‐diaminopropane), and [Zn(enMe)2‐(H2O)]4[Zn(enMe)2]2{(enMe)2{[Zn‐ (enMe)2]2[Zn4(HSiW9O34)2]}{[Zn‐ (enMe)2(H2O)]2[Zn4(HSiW9O34)2]}}? 13H2O ( 3 ) were hydrothermally synthesized and structurally characterized by elemental analysis, IR spectroscopy, thermogravimetric analysis, and single‐crystal X‐ray diffraction. Compound 1 consists of anions [Cu4(SiW9O34)2]12? linked by copper complexes into a 2D structure, whereas 2 is constructed from novel inorganic–organic hybrid anions [Zn4(HenMe)2(PW9O34)2]8? linked by zinc complexes into a 2D structure. The most interesting is the unique 2D network 3 , which consists of anions [Zn4(PW9O34)2]10? with two types of bridging groups: zinc complexes and enMe ligands.  相似文献   

8.
Two new heteropolyoxovanadoborates (H2dap)2H6{(VO)12O6[B3O6(OH)]6(H2O)}·13H2O (1, dap = 1,2-diaminopropane) and {[Zn(dien)]2[Zn(dien)(H2O)]4(VO)12O6[B3O6(OH)]6(H2O)}2·15H2O (2, dien = diethylenetriamine) have been hydrothermally synthesized and structurally characterized. Both 1 and 2 contain {(VO)12O6[B3O6(OH)]6(H2O)} cluster (denoted on V12B18), which is constructed by a puckered B18O36(OH)6 ring sandwiched between two triangles of six alternating cis and trans edge-sharing vanadium atoms, and a central water molecule. 1 consists of discrete [V12B18]10− cluster anions with H2dap2+ as counterions, while 2 consists of discrete neutral {[Zn(dien)]2[Zn(dien)(H2O)]4[V12B18]} clusters, which are built from two types of zinc(II) complex fragments connecting with V12B18 cluster through two Zn-(μ 3-O)-B bonds. Interestingly, 2 is the only example of the V12B18 cluster decorated by two types of zinc(II) complex fragments.  相似文献   

9.
Two new complexes of the Ln2(oda)3·nH2O (oda =–O2CCH2OCH2CO2–) series are reported, i.e. {[Pr2(C4H4O5)3(H2O)3]·5H2O}n and {[Nd2(C4H4O5)3(H2O)6]·C4H6O5·‐2H2O}n. The former is isostructural with the reported La analogue, while the latter is a new structural variety within the series. Each compound exhibits two independent nine‐coordinated Ln centres showing a variety of coordination geometries.  相似文献   

10.
Solvatothermal syntheses have been exploited to effect the isolation of three novel polyoxoalkoxometalate clusters, [{Fe(OH)(CH3CN)2} Fe6OCl6{(OCH2)3CCH2OH}4] (1), [Fe10O2Cl8{(OCH2)3CCH2CH3}6] (2), and [(VO)2Fe8O2Cl6{(OCH2)3CCH2CH3}6] (3). The structure of 1 may be described as a hexametalate core {Fe6OCl6}10+, consisting of a octahedral arrangement of chloride ligands encasing an octahedron of six Fe(III) sites, with a central oxo group. The remaining four coordination sites at each octahedral iron center are occupied by doubly bridging oxygen donors from the trisalkoxo ligands. One triangular face of this substructure, defined by three oxygen atoms, from three adjacent trisalkoxo ligands, is capped by the {Fe(OH)(CH3CN)2}2+ subunit. The structure of 2 is based on the decametalate core of edge-sharing octahedra. The eight peripheral Fe(III) sites of the cluster bond to four oxygen donors from the trisalkoxo ligands, a terminal Cl ligand, and one of the 6-oxo groups. The two central iron sites are linked to four oxygen donors from the trisalkoxo ligands and to both of the 6-oxo groups. Cluster 3 is structurally related to 2 with two {FeCl}2+ units replaced by {VO}2+ groups.  相似文献   

11.
The reactivity of [Rh(CO)2{(R,R)‐Ph? BPE}]BF4 ( 2 ) toward amine, CO and/or H2 was examined by high‐pressure NMR and IR spectroscopy. The two cationic pentacoordinated species [Rh(CO)3{(R,R)‐Ph? BPE}]BF4 ( 4 ) and [Rh(CO)2(NHC5H10){(R,R)‐Ph‐BPE}]BF4 ( 8 ) were identified. The transformation of 2 into the neutral complex [RhH(CO)2{(R,R)‐Ph? BPE}] ( 3 ) under hydroaminomethylation conditions (CO/H2, amine) was investigated. The full mechanisms related to the formation of 3 , 4 and 8 starting from 2 are supported by DFT calculations. In particular, the pathway from 2 to 3 revealed the deprotonation by the amine of the dihydride species [Rh(H)2(CO)2{(R,R)‐Ph? BPE}]BF4 ( 6 ), resulting from the oxidative addition of H2 on 2 .  相似文献   

12.
The tetranuclear compound [Mo2(O2C‐tBu)3]2(μ‐C2O4) ( 1 ) that is prepared from [Mo2(O2C‐tBu)3]4 and oxalic acid, was reacted with MnI2 · 2THF to form the polyoxomolybdate compound [Mn(CH3OH)6] [Mo8O16(OCH3)8(C2O4)] ( 2 ) in a complex redox reaction. Crystals of 2 were analyzed by single‐crystal X‐ray diffraction showing a octanuclear polyoxomolybdate dianion in which the Mo=O moieties are alternately connected through μ‐oxo and μ‐methoxo units. Charge balance in 2 is realized by a manganese(II) cation that is octahedrally coordinated by methanol ligands. The crystal structure is dominated by strong hydrogen bond interactions of the O–H ··· O type of methanol molecules coordinated to manganese as well as additional methanol molecules in the crystal lattice.  相似文献   

13.
A new series of ionic crystals, KH2[Cr3O(OOCCH3)6(H2O)3][α‐SiMo12O40] · 11 H2O ( 1 ), KH2[Cr3O(OOCCH3)6(H2O)3][α‐SiW12O40]μ·μ11H2O ( 2 ), K2[Cr3O(OOCCH3)6 (H2O)3][α‐PW12O40]μ· 17H2O ( 3 ), Na[Cr3O(OOCCH3)6(H2O)3]2[α‐PMo12O40] · 11H2O ( 4 ), H5[Cr3O(OOCCH3)6(H2O)3] [α‐P2Mo18O62] · 10H2O ( 5 ) based on a polyoxometalate building block with a macrocation, have been synthesized in aqueous solution and structurally characterized by single‐crystal X‐ray diffraction, IR spectra, elemental analysis, thermogravimetric analysis (TGA). The polyanions and macrocations stacked alternately through hydrogen bonds as well as electrostatic interactions to constitute a novel porous microstructure. In the crystal structures of 1 , 2 , and 3 , oppositely charged cluster ions stacked alternately to form one‐dimensional channels. Compound 4 exhibits an unique structure that six macrocation pillars packed along the a‐axis to form a straight 1D channel, in which accommodates a polyoxometalate pillar. In compound 5 , six α‐Dawson‐type polyoxometalate pillars stacked on top of each other along the a‐axis to form a straight 1D channel, which houses a macrocation pillar. The magnetic investigation on compounds 1 and 5 shows a typical antiferromagnetic interaction of the macrocation [Cr3O(OOCCH3)6(H2O)3]+, almost independent from the presence of polyoxometalate anions.  相似文献   

14.
A TG, DTG and DTA study of three polynuclear coordination compounds,containing Al(III)-Mg(II), namely (NH4)4[Al2Mg(C4O5H4)4(OH)4]?2H2O,(NH4)4[MgAl2(C4H4O6)4(OH)4]?3H2Oand (NH4)2[Al2Mg(C6O7H11)5(OH)5]?3H2O,has been reported together with the associated thermal decomposition mechanismrationalized in terms of intermediate products. As decomposition end-product,magnesium-aluminum spinel is obtained. The values of MgAl2O4mean crystallite size depend on the anionic ligand contained by the precursorcompound, varying in the order: malate (143 Å) ligand contained by theprecursor compound, varying in the order: malate (143 Å)  相似文献   

15.
Reaction of NiCl2 · 6H2O and P(CH2OH)3 (THP) with H2S and (H7O3)2[Mo6Cl14] · 3H2O in ethanol produces new trinuclear nickel sulphide complex [Ni33-S)2{(HOCH2)2PCH2OP(CH2OH)2}3][Mo6Cl14] · 0.8H2O (I) with new bidentate phosphine-phosphinite ligand resulted from THP condensation. It was characterized by X-ray structure analysis.  相似文献   

16.
An Anionic Oxohydroxo Complex with Bismuth(III): Na6[Bi2O2(OH)6](OH)2 · 4H2O Colourless, plate‐like, air sensitive crystals of Na6[Bi2O2(OH)6](OH)2 · 4H2O are obtained by reaction of Bi2O3 or Bi(NO3)3 · 5H2O in conc. NaOH (58 wt %) at 200 °C followed by slow cooling to room temperature. The crystal structure (triclinic, P 1¯, a = 684.0(2), b = 759.8(2), c = 822.7(2) pm, α = 92.45(3)°, ß = 90.40(3)°, γ = 115.60(2)°, Z = 1, R1, wR2 (all data), 0, 042, 0, 076) contains dimeric, anionic complexes [Bi2O2(OH)6]4— with bismuth in an ψ1‐octahedral coordination of two oxo‐ and three hydroxo‐ligands. The thermal decomposition was investigated by DSC/TG or DTA/TG and high temperature X‐ray powder diffraction measurements. In the final of three steps the decomposition product is Na3BiO3.  相似文献   

17.
The crystal structures of three unusual chromium organophosphate complexes have been determined, namely, bis(μ‐butyl 2,6‐di‐tert‐butyl‐4‐methylphenyl hydrogen phosphato‐κOO′)di‐μ‐hydroxido‐bis[(butyl 2,6‐di‐tert‐butyl‐4‐methylphenyl hydrogen phosphato‐κO)(butyl 2,6‐di‐tert‐butyl‐4‐methylphenyl phosphato‐κO)chromium](CrCr) heptane disolvate or {Cr22‐OH)22‐PO2(OBu)(O‐2,6‐tBu2‐4‐MeC6H2)‐κOO′]2[PO2(OBu)(O‐2,6‐tBu2‐4‐MeC6H2)‐κO]2[HOPO(OBu)(O‐2,6‐tBu2‐4‐MeC6H2)‐κO]2}·2C7H16, [Cr2(C19H32O4P)4(C19H33O4P)2(OH)2]·2C7H16, denoted ( 1 )·2(heptane), [μ‐bis(2,6‐diisopropylphenyl) phosphato‐1κO:2κO′]bis[bis(2,6‐diisopropylphenyl) phosphato]‐1κO,2κO‐chlorido‐2κCl‐triethanol‐1κ2O,2κO‐di‐μ‐ethanolato‐1κ2O:2κ2O‐dichromium(CrCr) ethanol monosolvate or {Cr22‐OEt)22‐PO2(O‐2,6‐iPr2‐C6H3)2‐κOO′][PO2(O‐2,6‐iPr2‐C6H3)2‐κO]2Cl(EtOH)3}·EtOH, [Cr2(C2H5O)2(C24H34O4P)3Cl(C2H6O)3]·C2H6O, denoted ( 2 )·EtOH, and di‐μ‐ethanolato‐1κ2O:2κ2O‐bis{[bis(2,6‐diisopropylphenyl) hydrogen phosphato‐κO][bis(2,6‐diisopropylphenyl) phosphato‐κO]chlorido(ethanol‐κO)chromium}(CrCr) benzene disolvate or {Cr22‐OEt)2[PO2(O‐2,6‐iPr2‐C6H3)2‐κO]2[HOPO(O‐2,6‐iPr2‐C6H3)2‐κO]2Cl2(EtOH)2}·2C6H6, [Cr2(C2H5O)2(C24H34O4P)2(C24H35O4P)2Cl2(C2H6O)2]·2C6H6, denoted ( 3 )·2C6H6. Complexes ( 1 )–( 3 ) have been synthesized by an exchange reaction between the in‐situ‐generated corresponding lithium or potassium disubstituted phosphates with CrCl3(H2O)6 in ethanol. The subsequent crystallization of ( 1 ) from heptane, ( 2 ) from ethanol and ( 3 ) from an ethanol/benzene mixture allowed us to obtain crystals of ( 1 )·2(heptane), ( 2 )·EtOH and ( 3 )·2C6H6, whose structures have the monoclinic P21, orthorhombic P212121 and triclinic P space groups, respectively. All three complexes have binuclear cores with a single Cr—Cr bond, i.e. Cr2O6P2 in ( 1 ), Cr2PO4 in ( 2 ) and Cr2O2 in ( 3 ), where the Cr atoms are in distorted octahedral environments, formally having 16 ē per Cr atom. The complexes have bridging ligands μ2‐OH in ( 1 ) or μ2‐OEt in ( 2 ) and ( 3 ). The organophosphate ligands demonstrate terminal κO coordination modes in ( 1 )–( 3 ) and bridging μ2‐κOO′ coordination modes in ( 1 ) and ( 2 ). All the complexes exhibit hydrogen bonding: two intramolecular Ophos…H—Ophos interactions in ( 1 ) and ( 3 ) form two {H[PO2(OR)2]2} associates; two intramolecular Cl…H—OEt hydrogen bonds additionally stabilize the Cr2O2 core in ( 3 ); two intramolecular Ophos…H—OEt interactions and two O…H—O intermolecular hydrogen bonds with a noncoordinating ethanol molecule are observed in ( 2 )·EtOH. The presence of both basic ligands (OH? or OEt?) and acidic [H(phosphate)2]? associates at the same metal centres in ( 1 ) and ( 3 ) is rather unusual. Complexes may serve as precatalysts for ethylene polymerization under mild conditions, providing polyethylene with a small amount of short‐chain branching. The formation of a small amount of α‐olefins has been detected in this reaction.  相似文献   

18.
The compounds (NMe4)5[As2Mo8V4AsO40] · 3 H2O 2a , (NH4)21[H3Mo57V6(NO)6O183(H2O)18] · 65 H2O 3a , (NH2Me2)18(NH4)6[Mo57V6(NO)6O183(H2O)18] · 14 H2O 3b and (NH4)12[Mo36(NO)4O108(H2O)16] · 33 H2O 4a ( 3a and 4a were not correctly reported in the literature regarding to their composition, structures and the oxidation states of the metal centres) which contain large isolated anionic species, have been prepared (among them 3a, 3b , and 4a in rather high yield) and characterized by complete crystal structure analysis as well as IR/Raman, UV/VIS/NIR, ESR spectroscopy and magnetic susceptibility measurements, redox titrations, bond valence sum calculations, elemental analyses and thermogravimetric studies. Perspectives for polyoxometalate chemistry referring to the synthesis of “extremely” large nanoscaled species are discussed, together with the occurrence of a large transferable {Mo17} building block in the compounds 3a, 3b and 4a which also exists in the corresponding iron compound Na3(NH4)12[H15Mo57Fe6(NO)6O183(H2O)18] · 76 H2O 7a .  相似文献   

19.
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   

20.
Synthesis, Characterization, and EPR Studies of Heteropoly Compounds with Iron(III) in Tetrahedral and Octahedral Coordination The heteropoly compounds H5[FeO4W12O36] · 6 H2O (a0 = 1216 pm), H3[Fe(OH)6Mo6O18] · 4 H2O, Na5[FeO4W12O36] · nH2O and FeH2[FeO4W12O36] · 17 H2O, for the first time obtained in this work by freeze-drying and characterized by means of chemical analysis, i.r. and u.v. spectroscopy, X-ray powder-photographs, and magnetic measurements, appear as suitable model systems for EPR investigations. They contain, like a number of known FeIII-heteropoly compounds, FeIII in FeO4 or/and FeO6 units, which are isolated from each other by structural reasons. In the Keggin-compounds M5[EIIIO4W12O36] · nH2O ( I ) (M = Na, Rb, TMA, TEA; E = Fe, Al, B) FeIII occupies slightly distorted tetrahedral positions (g′ ≈? 2), which are characterized by zfs-values of ≈? 10 mT and line widthes ΔB of 2.0 ?15 mT. Unlike as for I cations with different physico-chemical characteristics have only little effect on the FeIII-zfs. This holds for the Anderson-complexes M3[Fe(OH)6Mo6O18]·nH2O, (M = H, K, NH4, TMA; g′ ≈? 4.3 ΔB ≈? 67 mT) and for M5[SiO4W11O35FeO5(OH2)]·nH2O, (M = K, TMA; g′ = 4.3 ΔB = 26.5 mT). The FeO6 octahedra are more distorted than the FeO4 tetrahedra in I and therefore less susceptible for structural changes.  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