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 40 毫秒
1.
The reaction of bisdicyclohexylphosphinoethane (dcpe) and the subvalent MI sources [MI(PhF)2][pf] (M=Ga+, In+; [pf]=[Al(ORF)4]; RF=C(CF3)3) yielded the salts [{M(dcpe)}2][pf]2, containing the first dicationic, trans-bent digallene and diindene structures reported so far. The non-classical MI⇆MI double bonds are surprisingly short and display a ditetrylene-like structure. The bonding situation was extensively analyzed by quantum chemical calculations, QTAIM (Quantum Theory of Atoms in Molecules) and EDA-NOCV (Energy Decomposition Analysis with the combination of Natural Orbitals for Chemical Valence) analyses and is compared to that in the isoelectronic and isostructural, but neutral digermenes and distannenes. The dissolved [{Ga(dcpe)}2]2+([pf])2 readily reacts with 1-hexene, cyclooctyne, diphenyldisulfide, diphenylphosphine and under mild conditions at room temperature. This reactivity is analyzed and rationalized.  相似文献   

2.
Heterobimetallic cationic sandwich complexes [M(μ-Cp)M′Cp]+ of group 13 (M=Ga, In) and group 14 (M′=Ge, Sn) elements have been prepared as [WCA] salts (WCA=Al(ORF)4; ORF=OC(CF3)3). Their molecular structures include free apical gallium or indium atoms. The sandwich complexes were formed in the reactions of [M(HMB)]+[WCA] (HMB=C6Me6) with the free metallocenes [M′Cp2]. Their structures are related to known stannocene and stannocenium salts; the unprecedented germanium analogues, namely the free germanocenium cation [GeCp]+ and the corresponding triple-decker complex cation [CpGe(μ-Cp)GeCp]+, are described herein. By variation of the reaction conditions, these sandwich complexes can be transformed into the group 13/14 mixed cationic coordination polymer [{In(HMB)(μ-SnCp2)}n][WCA]n. This polymeric chain motif was also successfully replicated by the synthesis of complexes [{Ga/In(HMB)(μ-FeCp2)}n][WCA]n containing FeCp2 as a bridging ligand.  相似文献   

3.
Herein we report on the facile access to a GaI and InI complex salt of pentamethyldiethylenetriamine (PMDTA) with the weakly coordinating counterion [Al(ORF)4] [RF = C(CF3)3]. Thus, the low temperature reaction of [M(PhF)2]+ (M = Ga, In) solutions in 1,2-F2C6H4 (oDFB) with PMDTA yielded the title compounds M(PMDTA)+[Al(ORF)4] (M = Ga 1 , in 2 ) in non-optimized yields. Both compounds are highly sensitive towards air and moisture, have a rather limited lifetime at room temperature, but were structurally characterized.  相似文献   

4.
Several heterometallic nitrido complexes were prepared by reaction of the imido–nitrido titanium complex [{Ti(η5‐C5Me5)(μ‐NH)}33‐N)] ( 1 ) with amido derivatives of Group 13–15 elements. Treatment of 1 with bis(trimethylsilyl)amido [M{N(SiMe3)2}3] derivatives of aluminum, gallium, or indium in toluene at 150–190 °C affords the single‐cube amidoaluminum complex [{(Me3Si)2N}Al{(μ3‐N)23‐NH)Ti35‐C5Me5)33‐N)}] ( 2 ) or the corner‐shared double‐cube compounds [M(μ3‐N)33‐NH)3{Ti35‐C5Me5)33‐N)}2] [M=Ga ( 3 ), In ( 4 )]. Complexes 3 and 4 were also obtained by treatment of 1 with the trialkyl derivatives [M(CH2SiMe3)3] (M=Ga, In) at high temperatures. The analogous reaction of 1 with [{Ga(NMe2)3}2] at 110 °C leads to [{Ga(μ3‐N)23‐NH)Ti35‐C5Me5)33‐N)}2] ( 5 ), in which two [GaTi3N4] cube‐type moieties are linked through a gallium–gallium bond. Complex 1 reacts with one equivalent of germanium, tin, or lead bis(trimethylsilyl)amido derivatives [M{N(SiMe3)2}2] in toluene at room temperature to give cube‐type complexes [M{(μ3‐N)23‐NH)Ti35‐C5Me5)33‐N)}] [M=Ge ( 6 ), Sn ( 7 ), Pb ( 8 )]. Monitoring the reaction of 1 with [Sn{N(SiMe3)2}2] and [Sn(C5H5)2] by NMR spectroscopy allows the identification of intermediates [RSn{(μ3‐N)(μ3‐NH)2Ti35‐C5Me5)33‐N)}] [R=N(SiMe3)2 ( 9 ), C5H5 ( 10 )] in the formation of 7 . Addition of one equivalent of the metalloligand 1 to a solution of lead derivative 8 or the treatment of 1 with a half equivalent of [Pb{N(SiMe3)2}2] afford the corner‐shared double‐cube compound [Pb(μ3‐N)23‐NH)4{Ti35‐C5Me5)33‐N)}2] ( 11 ). Analogous antimony and bismuth derivatives [M(μ3‐N)33‐NH)3{Ti35‐C5Me5)33‐N)}2] [M=Sb ( 12 ), Bi ( 13 )] were obtained through the reaction of 1 with the tris(dimethylamido) reagents [M(NMe2)3]. Treatment of 1 with [AlCl2{N(SiMe3)2}(OEt2)] affords the precipitation of the singular aluminum–titanium square‐pyramidal aggregate [{{(Me3Si)2N}Cl3Al2}(μ3‐N)(μ3‐NH)2{Ti35‐C5Me5)3(μ‐Cl)(μ3‐N)}] ( 14 ). The X‐ray crystal structures of 5 , 11 , 13 , 14 , and [AlCl{N(SiMe3)2}2] were determined.  相似文献   

5.
Hydrothermally synthesized dipotassium gallium {hydrogen bis[hydrogenphosphate(V)]} difluoride, K2Ga[H(HPO4)2]F2, is isotypic with K2Fe[H(HPO4)2]F2. The main features of the structure are ([Ga{H(HPO4)2}F2]2−)n columns consisting of centrosymmetric Ga(F2O4) octahedra [average Ga—O = 1.966 (3) Å and Ga—F = 1.9076 (6) Å] stacked above two HPO4 tetrahedra [average P—O = 1.54 (2) Å] sharing two O‐atom vertices. The charge‐balancing seven‐coordinate K+ cations [average K—O,F = 2.76 (2) Å] lie in the intercolumn space, stabilizing a three‐dimensional structure. Strong [O...O = 2.4184 (11) Å] and medium [O...F = 2.6151 (10) Å] hydrogen bonds further reinforce the connections between adjacent columns.  相似文献   

6.
Digallane [L1Ga−GaL1] ( 1 , L1=dpp-bian=1,2-[(2,6-iPr2C6H3)NC]2C12H6) reacts with RN=C=O (R=Ph or Tos) by [2+4] cycloaddition of the isocyanate C=N bonds across both of its C=C−N−Ga fragments to afford [L1(O=C−NR)Ga−Ga(RN−C=O)L1] (R=Ph, 3 ; R=Tos, 4 ). The reactions with both isocyanates result in new C−C and N−Ga single bonds. In the case of allyl isocyanate, the [2+4] cycloaddition across one C=C−N−Ga fragment of 1 is accompanied by insertion of a second allyl isocyanate molecule into the Ga−N bond of the same fragment to afford compound [L1Ga−Ga(AllN− C=O)2L1] ( 5 ) (All=allyl). In the presence of Na metal, the related digallane [L2Ga−GaL2] ( 2 ; L2=dpp-dad=[(2,6-iPr2C6H3)NC(CH3)]2) is converted into the gallium(I) carbene analogue [L2Ga:] ( 2 A ), which undergoes a variety of reactions with isocyanate substrates. These include the cycloaddition of ethyl isocyanate to 2 A affording [Na2(THF)5]{L2Ga[EtN−C(O)]2GaL2} ( 6 ), cleavage of the N=C bond with release of 1 equiv. of CO to give [Na(THF)2]2[L2Ga(p-MeC6H4)(N−C(O))2−N(p-MeC6H4)]2 ( 7 ), cleavage of the C=O bond to yield the di-O-bridged digallium compound [Na(THF)3]2[L2Ga-(μ-O)2-GaL2] ( 8 ), and generation of the further addition product [Na2(THF)5][L2Ga(CyNCO2)]2 ( 9 ). Complexes 3 – 9 have been characterized by NMR (1H, 13C), IR spectroscopy, elemental analysis, and X-ray diffraction analysis. Their electronic structures have been examined by DFT calculations.  相似文献   

7.
Using [Ga(C6H5F)2]+[Al(ORF)4]?( 1 ) (RF=C(CF3)3) as starting material, we isolated bis‐ and tris‐η6‐coordinated gallium(I) arene complex salts of p‐xylene (1,4‐Me2C6H4), hexamethylbenzene (C6Me6), diphenylethane (PhC2H4Ph), and m‐terphenyl (1,3‐Ph2C6H4): [Ga(1,4‐Me2C6H4)2.5]+ ( 2+ ), [Ga(C6Me6)2]+ ( 3+ ), [Ga(PhC2H4Ph)]+ ( 4+ ) and [(C6H5F)Ga(μ‐1,3‐Ph2C6H4)2Ga(C6H5F)]2+ ( 52+ ). 4+ is the first structurally characterized ansa‐like bent sandwich chelate of univalent gallium and 52+ the first binuclear gallium(I) complex without a Ga?Ga bond. Beyond confirming the structural findings by multinuclear NMR spectroscopic investigations and density functional calculations (RI‐BP86/SV(P) level), [Ga(PhC2H4Ph)]+[Al(ORF)4]?( 4 ) and [(C6H5F)Ga(μ‐1,3‐Ph2C6H4)2Ga(C6H5F)]2+{[Al(ORF)4] ?}2 ( 5 ), featuring ansa‐arene ligands, were tested as catalysts for the synthesis of highly reactive polyisobutylene (HR‐PIB). In comparison to the recently published 1 and the [Ga(1,3,5‐Me3C6H3)2]+[Al(ORF)4]? salt ( 6 ) (1,3,5‐Me3C6H3=mesitylene), 4 and 5 gave slightly reduced reactivities. This allowed for favorably increased polymerization temperatures of up to +15 °C, while yielding HR‐PIB with high contents of terminal olefinic double bonds (α‐contents=84–93 %), low molecular weights (Mn=1000–3000 g mol?1) and good monomer conversions (up to 83 % in two hours). While the chelate complexes delivered more favorable results than 1 and 6 , the reaction kinetics resembled and thus concurred with the recently proposed coordinative polymerization mechanism.  相似文献   

8.
In a new oxidative route, Ag+[Al(ORF)4]? (RF=C(CF3)3) and metallic indium were sonicated in aromatic solvents, such as fluorobenzene (PhF), to give a precipitate of silver metal and highly soluble [In(PhF)n]+ salts (n=2, 3) with the weakly coordinating [Al(ORF)4]? anion in quantitative yield. The In+ salt and the known analogous Ga+[Al(ORF)4]? were used to synthesize a series of homoleptic PR3 phosphane complexes [M(PR3)n]+, that is, the weakly PPh3‐bridged [(Ph3P)3In–(PPh3)–In(PPh3)3]2+ that essentially contains two independent [In(PPh3)3]+ cations or, with increasing bulk of the phosphane, the carbene‐analogous [M(PtBu3)2]+ (M=Ga, In) cations. The MI? P distances are 27 to 29 pm longer for indium, and thus considerably longer than the difference between their tabulated radii (18 pm). The structure, formation, and frontier orbitals of these complexes were investigated by calculations at the BP86/SV(P), B3LYP/def2‐TZVPP, MP2/def2‐TZVPP, and SCS‐MP2/def2‐TZVPP levels.  相似文献   

9.
The homoleptic group 5 carbonylates [M(CO)6] (M=Nb, Ta) serve as ligands in carbonyl-terminated heterobimetallic AgmMn clusters containing 3 to 11 metal atoms. Based on our serendipitous [Ag6{Nb(CO)6}4]2+ ( 4 a 2+) precedent, we established access to such AgmMn clusters of the composition [Agm{M(CO)6}n]x (M=Nb, Ta; m=1, 2, 6; n=2, 3, 4, 5; x=1−, 1+, 2+). Salts of those molecular cluster ions were synthesized by the reaction of [NEt4][M(CO)6] and Ag[Al(ORF)4] (RF=C(CF3)3) in the correct stoichiometry in 1,2,3,4-tetrafluorobenzene at −35 °C. The solid-state structures were determined by single-crystal X-ray diffraction methods and, owing to the thermal instability of the clusters, a limited scope of spectroscopic methods. In addition, DFT-based AIM calculations were performed to provide an understanding of the bonding within these clusters. Apparently, the clusters 3 + (m=6, n=5) and 4 2+ (m=6, n=4) are superatom complexes with trigonal-prismatic or octahedral Ag6 superatom cores. The [M(CO)6] ions then bind through three CO units as tridentate chelate ligands to the superatom core, giving overall structures related to tetrahedral AX4 ( 4 2+) or trigonal bipyramidal AX5 molecules ( 3 +).  相似文献   

10.
New copper(II) complexes of the hydrazone ligands H2salhyhb, H2salhyhp, and H2salhyhh, derived from salicylaldehyde and ω‐hydroxy carbonic acid hydrazides, have been synthesized and physically characterized. Two fundamental structures were found in solid state depending on the pH‐value of the reaction solution. Acidic conditions lead to the formation of the di‐μ‐phenoxo‐bridged dicationic complex dimers [{Cu(Hsalhyhb)}2]2+ ( 1a ), [{Cu(Hsalhyhp)}2]2+ ( 2a ), and [{Cu(Hsalhyhh)}2]2+ ( 3a ), isolated as perchlorate salts. The dimeric complexes show strong antiferromagnetic coupling with J = ?399 ( 1a ), ?410 ( 2a ), and ?311 cm?1 ( 3a ). Higher pH‐values resulted in the aggregation of neutral copper ligand fragments to the one‐dimensional coordination polymers [{Cu(salhyhb)}n] ( 1b ), [{Cu(salhyhp)}n] ( 2b ), and [{Cu(salhyhh)}n] ( 3b ). 3b has been examined by means of X‐ray crystallography and represents the first example of a structurally characterized neutral copper(II) N‐salicylidenehydrazide complex without additional ligands. The magnetic interactions in the polymers are also antiferromagnetic with J = ?125 ( 1b ), ?136 ( 2b ), and ?148 cm?1 ( 3b ), but strongly reduced compared to the corresponding dimeric complexes. The two basic structure types can be reversibly interconverted simply by pH‐control.  相似文献   

11.
Subvalent Gallium Triflates – Potentially Useful Starting Materials for Gallium Cluster Compounds By reaction of GaCp* with trifluormethanesulfonic acid in hexane a mixture of gallium trifluormethanesulfonates (triflates, OTf) is obtained. This mixture reacts readily with lithiumsilanides [Li(thf)3Si(SiMe3)2R] (R = Me, SiMe3) to afford the cluster compounds [Ga6{Si(SiMe3)Me}6], [Ga2{Si(SiMe3)3}4] and [Ga10{Si(SiMe3)3}6]. By crystallization from various solvents the gallium triflates [Ga(OTf)3(thf)3], [HGa(OTf)(thf)4]+ [Ga(OTf)4(thf)3], [Cp*GaGa(OTf)2]2 and [Ga(toluene)2]+ [Ga5(OTf)6(Cp*)2] were isolated and characterized by single crystal X ray structure analysis.  相似文献   

12.
The reactions of the fluoride-ion donor, XeF6, with the fluoride-ion acceptors, M′OF4 (M′=Cr, Mo, W), yield [XeF5]+ and [Xe2F11]+ salts of [M′OF5] and [M2O2F9] (M=Mo, W). Xenon hexafluoride and MOF4 react in anhydrous hydrogen fluoride (aHF) to give equilibrium mixtures of [Xe2F11]+, [XeF5]+, [(HF)nF], [MOF5], and [M2O2F9] from which the title salts were crystallized. The [XeF5][CrOF5] and [Xe2F11][CrOF5] salts could not be formed from mixtures of CrOF4 and XeF6 in aHF at low temperature (LT) owing to the low fluoride-ion affinity of CrOF4, but yielded [XeF5][HF2]⋅CrOF4 instead. In contrast, MoOF4 and WOF4 are sufficiently Lewis acidic to abstract F ion from [(HF)nF] in aHF to give the [MOF5] and [M2O2F9] salts of [XeF5]+ and [Xe2F11]+. To circumvent [(HF)nF] formation, [Xe2F11][CrOF5] was synthesized at LT in CF2ClCF2Cl solvent. The salts were characterized by LT Raman spectroscopy and LT single-crystal X-ray diffraction, which provided the first X-ray crystal structure of the [CrOF5] anion and high-precision geometric parameters for [MOF5] and [M2O2F9]. Hydrolysis of [Xe2F11][WOF5] by water contaminant in HF solvent yielded [XeF5][WOF5]⋅XeOF4. Quantum-chemical calculations were carried out for M′OF4, [M′OF5], [M′2O2F9], {[Xe2F11][CrOF5]}2, [Xe2F11][MOF5], and {[XeF5][M2O2F9]}2 to obtain their gas-phase geometries and vibrational frequencies to aid in their vibrational mode assignments and to assess chemical bonding.  相似文献   

13.
Reactions of organomagnesium halides with group 13 metal halides lead to the formation of R3M type compounds (R = alkyl, aryl; M = Al, Ga, In) and are considered as the simplest methods of R3M compound syntheses. These seemingly simple reactions reveal a much more complex chemistry involving mixed magnesium-group 13 metal compounds. To elucidate the reaction course of reactions of organomagnesium halides with group 13 metal halides, we have studied reactions of R3M with organomagnesium halides. The interaction of Et3M with R1MgX led to the formation of following products being mixtures of crystalline ionic complexes with the general composition of [Et4-nR1nM][XMg (thf)5]+·(thf): [Et2.2Al(CH=CH2)1.8][BrMg (thf)5]+·(thf) ( 1 ), [Et3Ga(CH=CH2)][BrMg (thf)5]+·(thf) ( 2 ), [Et4Al][BrMg (thf)5]+·(thf) ( 3 ), [Et4Ga][BrMg (thf)5]+·(thf) ( 4 ), [Et2.9Al(C6H5)1.1][BrMg (thf)5]+·(thf) ( 5 ), [Et2.9Ga(C6H5)1.1][BrMg (thf)5]+·(thf) ( 6 ), [Et3.4GaMe0.6][IMg (thf)5]+·(thf) ( 7 ) and [Et4In][BrMg (thf)5]+·(thf) ( 8 ). A comparison of the production course of group 13 metal trialkyls R3M with a thermal decomposition of 1–8 products showed that reactions of MX3 with RMgX (X = Br, I; R = alkyl, aryl) yield initially intermediate ionic compounds, which must then be thermally decomposed to obtain pure R3M compounds. If group 13 metal bromides and iodides, and alkyl (aryl)magnesium bromides and iodides in thf are used, only intermediate products with the [R4M][XMg (thf)5]+·(thf) structure are formed.  相似文献   

14.
The syntheses of the two novel complexes [Ag{Mo/W(CO)6}2]+[F-{Al(ORF)3}2] (RF=C(CF3)3) are reported along with their structural and spectroscopic characterization. The X-ray structure shows that three carbonyl ligands from each M(CO)6 fragment bend towards the silver atom within binding Ag−C distance range. DFT calculations of the free cations [Ag{M(CO)6}2]+ (M=Cr, Mo, W) in the electronic singlet state give equilibrium structures with C2 symmetry with two bridging carbonyl groups from each hexacarbonyl ligand. Similar structures with C2 symmetry (M=Nb) and D2 symmetry (M=V, Ta) are calculated for the isoelectronic group 5 anions [Ag{M(CO)6}2] (M=V, Nb, Ta). The electronic structure of the cations is analyzed with the QTAIM and EDA-NOCV methods, which provide detailed information about the nature of the chemical bonds between Ag+ and the {M(CO)6}2q (q = −2, M = V, Nb, Ta; q = 0, M = Cr, Mo, W) ligands.  相似文献   

15.
The reaction of fumaryl fluoride with the superacidic solutions XF/MF5 (X=H, D; M=As, Sb) results in the formation of the monoprotonated and diprotonated species, dependent on the stoichiometric ratio of the Lewis acid to fumaryl fluoride. The salts [C4H3F2O2]+[MF6] (M=As, Sb) and [C4H2X2F2O2]2+([MF6])2 (X=H, D; M=As, Sb) are the first examples with a protonated acyl fluoride moiety. They were characterized by low-temperature vibrational spectroscopy. Low-temperature NMR spectroscopy and single-crystal X-ray structure analyses were carried out for [C4H3F2O2]+[SbF6] as well as for [C4H4F2O2]2+([MF6])2 (M=As, Sb). The experimental results are discussed together with quantum chemical calculations of the cations [C4H4F2O2 ⋅ 2 HF]2+ and [C4H3F2O2 ⋅ HF]+ at the B3LYP/aug-cc-pVTZ level of theory. In addition, electrostatic potential (ESP) maps combined with natural population analysis (NPA) charges were calculated in order to investigate the electron distribution and the charge-related properties of the diprotonated species. The C−F bond lengths in the protonated dication are considerably reduced on account of the +R effect.  相似文献   

16.
UO2F2 abstracts F anions from TlF in liquid ammonia solution and the compound [Tl2(NH3)6][{UO2F2(NH3)}2(μ-F)2] is formed. The compound has been characterized by single crystal X-ray diffraction, Raman spectroscopy and quantum-chemical calculations for the solid state. Quantum-chemical investigation of the [{UO2F2(NH3)}2(μ-F)2]2− anion showed that the U−(μ-F)−U σ-3c-4e-bond is essentially ionic. The [Tl2(NH3)6]2+ cation shows a thallophilic Tl⋅⋅⋅Tl interaction. Fluoride ion affinities (FIAs) were calculated for different UO22+ species [UO2Fx]2−x and [UO2Fx(NH3)5−x]2−x with x=0 to 4.  相似文献   

17.
The reduction of digallane [(dpp‐bian)Ga? Ga(dpp‐bian)] ( 1 ) (dpp‐bian=1,2‐bis[(2,6‐diisopropylphenyl)imino]acenaphthene) with lithium and sodium in diethyl ether, or with potassium in THF affords compounds featuring the direct alkali metal–gallium bonds, [(dpp‐bian)Ga? Li(Et2O)3] ( 2 ), [(dpp‐bian)Ga? Na(Et2O)3] ( 3 ), and [(dpp‐bian)Ga? K(thf)5] ( 7 ), respectively. Crystallization of 3 from DME produces compound [(dpp‐bian)Ga? Na(dme)2] ( 4 ). Dissolution of 3 in THF and subsequent crystallization from diethyl ether gives [(dpp‐bian)Ga? Na(thf)3(Et2O)] ( 5 ). Ionic [(dpp‐bian)Ga]?[Na([18]crown‐6)(thf)2]+ ( 6 a ) and [(dpp‐bian)Ga]?[Na(Ph3PO)3(thf)]+ ( 6 b ) were obtained from THF after treatment of 3 with [18]crown‐6 and Ph3PO, respectively. The reduction of 1 with Group 2 metals in THF affords [(dpp‐bian)Ga]2M(thf)n (M=Mg ( 8 ), n=3; M=Ca ( 9 ), Sr ( 10 ), n=4; M=Ba ( 11 ), n=5). The molecular structures of 4 – 7 and 11 have been determined by X‐ray crystallography. The Ga? Na bond lengths in 3 – 5 vary notably depending on the coordination environment of the sodium atom.  相似文献   

18.
By reaction of tmp2GaCl (tmp = 2,2,6,6‐tetramethylpiperidino) ( 1 ) with dilithioferrocene the 1,1′‐digallylferrocene [{Fe(η5‐C5H4)2}(tmp2Ga)2] ( 2 ) was prepared. 2 reacted with CO2 to afford the ferrocenophane [Fe{η5‐C5H4‐Ga(O2Ctmp)(μ2‐O2Ctmp)}2] ( 3 ). Here, an eight‐membered Ga(OCO)2Ga‐Ring is the bridge in the ferrocenophane structure. The gallium carbamates [Me2GaO2Ctmp]2 ( 5 ), and tmp2Ga(η2‐O2Ctmp) ( 7 ) show structural features embedded in 3 . The compounds were fully characterized by NMR, cyclovoltammetry and single crystal X‐ray structure analysis.  相似文献   

19.
Hydrothermal reactions of V2O5, tetra-2-pyridylpyrazine (tpyprz) and an appropriate M(II) starting material yield a series of oxides of general composition [{Mx(tpyprz)}yV4O12] [x=1, y=2, M=Co(II), Ni(II); x=2, y=2, M=Cu(I); x=2, y=1, M=Zn(II)]. The Co(II) and Ni(II) analogues (1 and 2) are isostructural and consist of one-dimensional ribbons constructed from {V4O12}4− clusters linked through {M(tpyprz)}24+ binuclear units of edge sharing {MO3N3} octahedra. In contrast, the structure of [{Cu2(tpyprz)}2V4O12] (3) is two-dimensional and constructed of {Cu2(tpyprz)}n2n+ chains linked in the second dimension through the {V4O12}4− clusters. The structure of [{Zn2(tpyprz)V4O12] (4) is also two-dimensional but may be described as {Zn2V4O12} chains interconnected through the binucleating tpyprz ligands. The roles of the coordination preferences of the secondary metal cations as well as the nature of the organic components are discussed.  相似文献   

20.
The complexes Ag(L)n[WCA] (L=P4S3, P4Se3, As4S3, and As4S4; [WCA]=[Al(ORF)4] and [F{Al(ORF)3}2]; RF=C(CF3)3; WCA=weakly coordinating anion) were tested for their performance as ligand-transfer reagents to transfer the poorly soluble nortricyclane cages P4S3, P4Se3, and As4S3 as well as realgar As4S4 to different transition-metal fragments. As4S4 and As4S3 with the poorest solubility did not yield complexes. However, the more soluble silver-coordinated P4S3 and P4Se3 cages were transferred to the electron-poor Fp+ moiety ([CpFe(CO)2]+). Thus, reaction of the silver salt in the presence of the ligand with Fp−Br yielded [Fp−P4S3][Al(ORF)4] ( 1 a ), [Fp−P4S3][F(Al(ORF)3)2] ( 1 b ), and [Fp−P4Se3][Al(ORF)4] ( 2 ). Reactions with P4S3 also yielded [FpPPh3−P4S3][Al(ORF)4] ( 3 ), a complex with the more electron-rich monophosphine-substituted Fp+ analogue [FpPPh3]+ ([CpFe(PPh3)(CO)]+). All complex salts were characterized by single-crystal XRD, NMR, Raman, and IR spectroscopy. Interestingly, they show characteristic blueshifts of the vibrational modes of the cage, as well as structural contractions of the cages upon coordination to the Fp/FpPPh3 moieties, which oppose the typically observed cage expansions that lead to redshifts in the spectra. Structure, bonding, and thermodynamics were investigated by DFT calculations, which support the observed cage contractions. Its reason is assigned to σ and π donation from the slightly P−P and P−E antibonding P4E3-cage HOMO (e symmetry) to the metal acceptor fragment.  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