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.
This paper overviews three living cationic polymerization systems (for styrene, p-methoxystyrene, and isobutyl vinyl ether) that are, in common, featured by: (i) specifically in nonpolar solvents, the use of the hydrogen halide/metal halide initiating systems (HX/MXn; X: I, Br, Cl; MXn: ZnX2, SnCl4), which generate a living growing carbocation stabilized by a nucleophilic counteranion (X…MXn); (ii) specifically in polar solvents, the use of externally added ammonium salts (nBu4N+Y; Y: I, Br, Cl), which permit the generation of living species from HX/MXn by providing nucleophilic halogen anions Y, either the same as or different from the halogen X in HX.  相似文献   

2.
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   

3.
The title salt, [Zn(C2N2H8)3]2[CdI4]I2, conventionally abbreviated [Zn(en)3]2[CdI4]I2, where en is ethyl­enediamine, contains discrete [Zn(en)3]2+ cations and [CdI4]2− anions with distorted octa­hedral and nearly tetra­hedral geometries, respectively, as well as uncoordinated I ions. The cation and the free I anion lie on twofold rotation axes and the [CdI4]2− anion lies on a axis in the space group I2d. The structure exhibits numerous weak inter‐ionic hydrogen bonds of two types, viz. N—H⋯I(free ion) and N—H⋯I([CdI4]2−), which support the resulting three‐dimensional framework.  相似文献   

4.
A series of octanuclear iodine-bromine interhalides [InBr8−n]2− (n=0, 2, 3, 4) were prepared systematically in two steps. Firstly, addition of a dihalogen (Br2 or IBr) to the triaminocyclopropenium bromide salt [C3(NEt2)3]Br forms the corresponding trihalide salt with Br3 or IBr2 anions, respectively. Secondly, addition to Br3 of half an equivalent of Br2 gives the octabromine polyhalide [Br8]2−, whereas addition to IBr2 of half an equivalent of Br2, IBr or I2 gives the corresponding interhalides: [I2Br6]2−, [I3Br5]2−, and [I4Br4]2−, respectively. The four octahalides were characterized by X-ray crystallography, computational studies, Raman and Far-IR spectroscopies, as well as by TGA and melting point. All of the salts were found to be ionic liquids.  相似文献   

5.
The trapping of a silicon(I) radical with N-heterocyclic carbenes is described. The reaction of the cyclic (alkyl)(amino) carbene [cAACMe] (cAACMe=:C(CMe2)2(CH2)NAr, Ar=2,6-iPr2C6H3) with H2SiI2 in a 3:1 molar ratio in DME afforded a mixture of the separated ion pair [(cAACMe)2Si:.]+I ( 1 ), which features a cationic cAAC–silicon(I) radical, and [cAACMe−H]+I. In addition, the reaction of the NHC–iodosilicon(I) dimer [IAr(I)Si:]2 (IAr=:C{N(Ar)CH}2) with 4 equiv of IMe (:C{N(Me)CMe}2), which proceeded through the formation of a silicon(I) radical intermediate, afforded [(IMe)2SiH]+I ( 2 ) comprising the first NHC–parent-silyliumylidene cation. Its further reaction with fluorobenzene afforded the CAr−H bond activation product [1-F-2-IMe-C6H4]+I ( 3 ). The isolation of 2 and 3 confirmed the reaction mechanism for the formation of 1 . Compounds 1 – 3 were analyzed by EPR and NMR spectroscopy, DFT calculations, and X-ray crystallography.  相似文献   

6.
The trapping of a silicon(I) radical with N‐heterocyclic carbenes is described. The reaction of the cyclic (alkyl)(amino) carbene [cAACMe] (cAACMe=:C(CMe2)2(CH2)NAr, Ar=2,6‐i Pr2C6H3) with H2SiI2 in a 3:1 molar ratio in DME afforded a mixture of the separated ion pair [(cAACMe)2Si:.]+I ( 1 ), which features a cationic cAAC–silicon(I) radical, and [cAACMe−H]+I. In addition, the reaction of the NHC–iodosilicon(I) dimer [IAr(I)Si:]2 (IAr=:C{N(Ar)CH}2) with 4 equiv of IMe (:C{N(Me)CMe}2), which proceeded through the formation of a silicon(I) radical intermediate, afforded [(IMe)2SiH]+I ( 2 ) comprising the first NHC–parent‐silyliumylidene cation. Its further reaction with fluorobenzene afforded the CAr−H bond activation product [1‐F‐2‐IMe‐C6H4]+I ( 3 ). The isolation of 2 and 3 confirmed the reaction mechanism for the formation of 1 . Compounds 1 – 3 were analyzed by EPR and NMR spectroscopy, DFT calculations, and X‐ray crystallography.  相似文献   

7.
Heterometallic metal–organic frameworks (MOFs) allow the precise placement of various metals at atomic precision within a porous framework. This new level of control by MOFs promises fascinating advances in basic science and application. However, the rational design and synthesis of heterometallic MOFs remains a challenge due to the complexity of the heterometallic systems. Herein, we show that bimetallic MOFs with MX2(INA)4 moieties (INA=isonicotinate; M=Co2+ or Fe2+; X=OH?, Cl?, Br?, I?, NCS?, or NCSe?) can be generated by the sequential modification of a Zr‐based MOF. This multi‐step modification not only replaced the linear organic linker with a square planar MX2(INA)4 unit, but also altered the symmetry, unit cell, and topology of the parent structure. Single‐crystal to single‐crystal transformation is realized so that snapshots for transition process were captured by successive single‐crystal X‐ray diffraction. Furthermore, the installation of Co(NCS)2(INA)4 endows field‐induced slow magnetic relaxation property to the diamagnetic Zr‐MOF.  相似文献   

8.
The synergistic Ag+/X2 system (X=Cl, Br, I) is a very strong, but ill‐defined oxidant—more powerful than X2 or Ag+ alone. Intermediates for its action may include [Agm(X2)n]m+ complexes. Here, we report on an unexpectedly variable coordination chemistry of diiodine towards this direction: ( A )Ag‐I2‐Ag( A ), [Ag2(I2)4]2+( A )2 and [Ag2(I2)6]2+( A )2⋅(I2)x≈0.65 form by reaction of Ag( A ) ( A =Al(ORF)4; RF=C(CF3)3) with diiodine (single crystal/powder XRD, Raman spectra and quantum‐mechanical calculations). The molecular ( A )Ag‐I2‐Ag( A ) is ideally set up to act as a 2 e oxidant with stoichiometric formation of 2 AgI and 2 A . Preliminary reactivity tests proved this ( A )Ag‐I2‐Ag( A ) starting material to oxidize n‐C5H12, C3H8, CH2Cl2, P4 or S8 at room temperature. A rough estimate of its electron affinity places it amongst very strong oxidizers like MF6 (M=4d metals). This suggests that ( A )Ag‐I2‐Ag( A ) will serve as an easily in bulk accessible, well‐defined, and very potent oxidant with multiple applications.  相似文献   

9.

The reaction of cationic clusters [Mo3S4(Dppet)3X3]X (I; Dppet = cis-1,2-bis(diphenylphosphino)ethylene; X = Cl, Br) with gallium metal in tetrahydrofuran (THF) affords cubane-like clusters [Mo3S4(GaX)(Dppet)3X3] (IIa, X = Cl; IIb, X = Br). In the synthesis of IIb, the crystals of the cationic cluster [Mo3S4(Dppet)3Br3](GaBr4) (III) were isolated from the mother liquor. The structures of complexes IIa ? 1.5THF, IIb ? 1.5THF, III ? 1.625THF, and [Mo3S4(Dppe)3Br3](GaBr4) ? 2THF (IV ? 2THF; Dppe = 1,2-bis(diphenylphosphino)ethane) were established by X-ray diffraction (CIF files CCDC № 1851341–1851344).

  相似文献   

10.
The reactions of three different tetracoordinated Ir complexes, [Ir(troppph)2]n (n=+1, 0, −1), which differ in the formal oxidation state of the metal from +1 to −1, with proton sources and dihydrogen were investigated (tropp=5‐(diphenylphosphanyl)dibenzo[a,d]cycloheptene). It was found that the cationic 16‐electron complex [Ir(troppph)2]+ ( 2 ) cannot be protonated but reacts with NaBH4 to the very stable 18‐electron IrI hydride [IrH(troppph)2] ( 5 ), which is further protonated with medium strong acids to give the 18‐electron IrIII dihydride [IrH2(troppph)2]+ ( 6 ; pKs in CH2Cl2/THF/H2O 1 : 1 : 2 ca. 2.2). Both, the neutral 17‐electron Ir0 complex [Ir(troppph)2] ( 3 ) and the anionic 18‐electron complex [Ir(troppph)2] ( 4 ) react rapidly with H2O to give the monohydride 5 . In reactions of 3 with H2O, the terminal IrI hydroxide [Ir(OH)(troppph)2] ( 8 ) is formed in equal amounts. All these complexes, apart from 5 , which is inert, do react rapidly with dihydrogen. The complex 2 gives the dihydride 6 in an oxidative addition reaction, while 3 , 4 , and 8 give the monohydride 5 . Interestingly, a salt‐type hydride (i.e., LiH) is formed as further product in the unexpected reaction with [Li(thf)x]+[Ir(troppph)2] ( 4 ). Because 3 undergoes disproportionation into 2 and 4 according to 2 3 ⇄ 2 + 4 (Kdisp=2.7⋅10−5), it is likely that actually the diamagnetic species and not the odd‐electron complex 3 is involved in the reactions studied here, and possible mechanisms for these are discussed.  相似文献   

11.
The straightforward synthesis of the cationic, purely organometallic NiI salt [Ni(cod)2]+[Al(ORF)4] was realized through a reaction between [Ni(cod)2] and Ag[Al(ORF)4] (cod=1,5‐cyclooctadiene). Crystal‐structure analysis and EPR, XANES, and cyclic voltammetry studies confirmed the presence of a homoleptic NiI olefin complex. Weak interactions between the metal center, the ligands, and the anion provide a good starting material for further cationic NiI complexes.  相似文献   

12.
The three-coordinate aluminum cations ligated by N-heterocyclic carbenes (NHCs) [(NHC) ⋅ AlMes2]+[B(C6F5)4] (NHC=IMeMe 4 , IiPrMe 5 , IiPr 6 , Mes=2,4,6-trimethylphenyl) were prepared via hydride abstraction of the alanes (NHC) ⋅ AlHMes2 (NHC=IMeMe 1 , IiPrMe 2 , IiPr 3 ) using [Ph3C]+[B(C6F5)4] in toluene as hydride acceptor. If this reaction was performed in diethyl ether, the corresponding four-coordinate aluminum etherate cations [(NHC) ⋅ AlMes2(OEt2)]+ [B(C6F5)4] 7 – 9 (NHC=IMeMe 7 , IiPrMe 8 , IiPr 9 ) were isolated. According to a theoretical and experimental assessment of the Lewis-acidity of the [(IMeMe) ⋅ AlMes2]+ cation is the acidity larger than that of B(C6F5)3 and of similar magnitude as reported for Al(C6F5)3. The reaction of [(IMeMe) ⋅ AlMes2]+[B(C6F5)4] 4 with the sterically less demanding, basic phosphine PMe3 afforded a mixed NHC/phosphine stabilized cation [(IMeMe) ⋅ AlMes2(PMe3)]+[B(C6F5)4] 10 . Equimolar mixtures of 4 and the sterically more demanding PCy3 gave a frustrated Lewis-pair (FLP), i.e., [(IMeMe) ⋅ AlMes2]+[B(C6F5)4]/PCy3 FLP-11 , which reacts with small molecules such as CO2, ethene, and 2-butyne.  相似文献   

13.
The cations [Pd 2 Cl 2 L] 2+ and [KL 2 + (L = [18]aneN2S4, L′ =[15]aneO5) have been used as templates for the synthesis of unique three-dimensional polyiodide networks. The metal cations in [Pd2Cl2L]1.5I5(I3)2 are linked into infinite chains by pairwise hydrogen bonding; the resulting cationic polymers run through channels formed by the extended polyiodide network. [KL2]I9 shows a three-dimensional network of puckered cubic cages of I9 ions whose cavities are occupied by the metal cations (section from the structure shown on the right).  相似文献   

14.
The synthesis of Naumann's AgI/AgIII mixed valence salt [AgI]+[AgIII(CF3)4] ( Ag-1 ) is revisited. Ag-1 is now safely available in half gram scale upon 2e oxidation of AgF in presence of CF3SiMe3 and ambient air. In addition to its unprecedented crystallographic characterization, the use of Ag-1 to build the novel AgI/AgIII salts [ Ag (bpy)2] -1 , [ Ag (18-crown-6)2] -1 , [ Ag -crypt-222] -1 and [ Ag (PCy3)2] -1 is herein reported, alongside their characterization by NMR, single crystal X-ray diffraction (Sc-XRD) and elemental analysis (EA). The utility of the currently affordable Ag-1 in gold(I) catalysis was demonstrated by the excellent catalytic activity displayed by [{ Au (PPh3)}2(μ-Cl)] -1 and [ Au (PPh3)] -1 in the 5-exo-dig cyclization of N-propargylbenzamide ( 2 ). These cationic AuI catalysts are accessible from (PPh3)AuCl and Ag-1 , and outperform the activity of the well-known benchmark catalyst (PPh3)AuNTf2.  相似文献   

15.
The title compound, {[Cu(NH3)4][Cu(CN)3]2}n, features a CuI–CuII mixed‐valence CuCN framework based on {[Cu2(CN)3]}n anionic layers and [Cu(NH3)4]2+ cations. The asymmetric unit contains two different CuI ions and one CuII ion which lies on a centre of inversion. Each CuI ion is coordinated to three cyanide ligands with a distorted trigonal–planar geometry, while the CuII ion is ligated by four ammine ligands, with a distorted square‐planar coordination geometry. The interlinkage between CuI ions and cyanide bridges produces a honeycomb‐like {[Cu2(CN)3]}n anionic layer containing 18‐membered planar [Cu(CN)]6 metallocycles. A [Cu(NH3)4]2+ cation fills each metallocyclic cavity within pairs of exactly superimposed {[Cu2(CN)3]}n anionic layers, but there are no cations between the layers of adjacent pairs, which are offset. Pairs of N—H...N hydrogen‐bonding interactions link the N—H groups of the ammine ligands to the N atoms of cyanide ligands.  相似文献   

16.
Treatment of the coordinatively unsaturated cationic complexes, [(η-C3H5)(Ph3P)2PdII]PF6 and [(1,5-COD)2RhI]BF4, with potassium t-butylperoxide in dichloromethane gives the t-butylperoxometal complexes; (η-C3H5)(Ph3P)(t-BuOO)PdII and [(1,5-COD)(t-BuOO)RhI](KBF4), via nucleophilic attack by t-BuOO on the cationic metal center.  相似文献   

17.
We report on the synthesis of a new metal–organic framework (MOF) composed of Sn(OCH3)2–tetrakis(pyridin‐4‐yl)porphyrin linkers, Cu+ connecting nodes and [CuCl2] counter‐ions, namely poly[[bis(methanolato‐κO)[μ5‐5,10,15,20‐tetrakis(pyridin‐4‐yl)porphyrin‐κ8N5:1′κN10:1′′κN15:1′′′κN20:2κ4N21,N22,N23,N24]copper(I)tin(II)] dichloridocuprate(I)], [CuSn(C40H24N8)(CH3O)2][CuCl2]. Its crystal structure consists of a single‐framework coordination polymer of the organic ligand and the CuI ions. The latter are characterized by a tetrahedral coordination geometry [with CN (coordination number) = 4], linking to the pyridyl N‐atom sites of four different ligands and imparting to the positively charged polymeric assembly a diamondoid PtS‐type topology. Correspondingly, every porphyrin unit is coordinated to four different CuI connectors. The [CuCl2] anions occupy the intra‐lattice voids, along with disordered molecules of the water crystallization solvent. The asymmetric unit of this structure consists of two halves of the porphyrin scaffold, located on centres of crystallographic inversion, and the Cu+ and [CuCl2] ions. This report provides unique structural evidence for the formation of tetrapyridylporphyrin‐based three‐dimensional MOFs with a diamondoid architecture that have been observed earlier only on rare occasions.  相似文献   

18.
The reaction of Na[CoIII(d -ebp)] (d -H4ebp = N,N′-ethylenebis[d -penicillamine]) with [(AuICl)2(dppe)] (dppe = 1,2-bis[diphenylphosphino]ethane) gave a cationic AuI4CoIII2 hexanuclear complex, [CoIII2(LAu4)]2+ ([ 1 ]2+), where [LAu4]4− is a cyclic tetragold(I) metalloligand with a 32-membered ring, [AuI4(dppe)2(d -ebp)2]4−. Complex [ 1 ]2+ crystallized with NO3 to produce a charge-separation (CS)-type ionic solid of [ 1 ](NO3)2. In [ 1 ](NO3)2, the complex cations are assembled to form cationic supramolecular hexamers of {[ 1 ]2+}6, which are closely packed in a face-centered cubic (fcc) lattice structure. The nitrate anions of [ 1 ](NO3)2 were accommodated in hydrophilic and hydrophobic tetrahedral interstices of the fcc structure to form tetrameric and hexameric nitrate clusters of {NO3}4 and {NO3}6, respectively. An analogous CS-type ionic solid formulated as [NiIICoIII(LAu4)](NO3) ([ 2 ](NO3)) was obtained when a 1:1 mixture of Na[CoIII(d -ebp)] and [NiII(d -H2ebp)] was reacted with [(AuICl)2(dppe)], accompanied by the conversion of the diamagnetic, square-planar [NiII(d -H2ebp)] to the paramagnetic, octahedral [NiII(d -ebp)]2−. While the overall fcc structure in [ 2 ](NO3) was similar to that of [ 1 ](NO3)2, none of the nitrate anions were accommodated in any hydrophobic tetrahedral interstice, reflecting the difference in the complex charges between [ 1 ]2+ and [ 2 ]+.  相似文献   

19.
The propulsion of photocatalytic hydrogen (H2) production is limited by the rational design and regulation of catalysts with precise structures and excellent activities. In this work, the [MoOS3]2− unit is introduced into the CuI clusters to form a series of atomically-precise MoVI−CuI bimetallic clusters of [Cu6(MoOS3)2(C6H5(CH2)S)2(P(C6H4R)3)4] ⋅ xCH3CN (R=H, CH3, or F), which show high photocatalytic H2 evolution activities and excellent stability. By electron push-pull effects of the surface ligand, highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) levels of these MoVI−CuI clusters can be finely tuned, promoting the resultant visible-light-driven H2 evolution performance. Furthermore, MoVI−CuI clusters loaded onto the surface of magnetic Fe3O4 carriers significantly reduced the loss of catalysts in the collection process, efficiently addressing the recycling issues of such small cluster-based catalyst. This work not only highlights a competitively universal approach on the design of high-efficiency cluster photocatalysts for energy conversion, but also makes it feasible to manipulate the catalytic performance of clusters through a rational substituent strategy.  相似文献   

20.
We present a crystal engineering strategy to fine tune the pore chemistry and CH4‐storage performance of a family of isomorphic MOFs based upon PCN‐14. These MOFs exhibit similar pore size, pore surface, and surface area (around 3000 m2 g−1) and were prepared with the goal to enhance CH4 working capacity. [Cu2(L2)(H2O)2]n (NJU‐Bai 41: NJU‐Bai for Nanjing University Bai's group), [Cu2(L3)(H2O)2]n (NJU‐Bai 42), and [Cu2(L4)(DMF)2]n (NJU‐Bai 43) were prepared and we observed that the CH4 volumetric working capacity and volumetric uptake values are influenced by subtle changes in structure and chemistry. In particular, the CH4 working capacity of NJU‐Bai 43 reaches 198 cm3 (STP: 273.15 K, 1 atm) cm−3 at 298 K and 65 bar, which is amongst the highest reported for MOFs under these conditions and is much higher than the corresponding value for PCN‐14 (157 cm3 (STP) cm−3).  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