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 78 毫秒
1.
The preparation of a series of titanium half-sandwich compounds [Ti(η5-C5H5−x (SiMe3) x R3] (x = 1–3, R = Cl, Me) and their reactivity for propene polymerization is reported. The compounds 1–3 polymerize propene, albeit in a much lower activity than the reported [Ti(η5-C5Me5Me3]/B(C6F5)3 catalyst. Unlike the reported [Ti(η5-C5Me5Me3]/B(C6F5)3 catalyst, the quasi living polymerization was not observed. Instead, we observe rather unusual temperature effects when the trityl salt [Ph3C][B(C6F5)4] was used as activator. The activity increases with increasing temperature, whereas when B(C6F5)3 is used a decrease is observed The rather broad (>2) PDI indicates multisite catalysts, and 13C-NMR indicates predominantly atactic polypropene. The solid state structure of the hydrolysis product [{Ti(η5-C5H4(SiMe3)Cl2}O] (4) was determined.  相似文献   

2.
The voltammetric oxidation at a glassy carbon electrode of a series of ferrocenyl polyesters PmF{X}, (X=T, terephthalate; N, 2,6-naphthalene dicarboxylate; B, 4,4′-biphenyl dicarboxylate) is reported in the weakly donating and low-polarity CH2Cl2 solvent containing [NBu4][PF6], [NBu4][B(C6F5)4], or 1-butyl-1-methylpyrrolidinium bis(trifluoromethylsulfonyl)amide ([bmpyr][TFSA]) as the supporting electrolyte. The one electron oxidation of the ferrocenyl groups is strongly influenced by the nature of the supporting electrolyte anion. Use of the conventional [PF6] anion or the room temperature ionic liquid (RTIL), [bmpyr][TFSA] as the supporting electrolyte, gives rise to significant oxidation product interaction (precipitation and/or adsorption) with the electrode surface. In marked contrast, diffusion-controlled, chemically and electrochemically reversible processes are observed when the weakly coordinating [B(C6F5)4] is used as the anionic component of the supporting electrolyte. In this case, data obtained via cyclic voltammetry, chronoamperometry and chronocoulometry are consistent with ideal reversible one electron oxidation processes and a soluble cationic product. Diffusion coefficients of the monomers, polymers, ferrocene and decamethyferrocene are reported in the CH2Cl2/[NBu4][B(C6F5)4] system. Importantly, reversible potentials that are easily obtained under these conditions show that the acyl or methyl linkers, through which the ferrocenyl group is attached to the neo-pentylenediol component, tune the redox properties of the polymers. Electrochemical studies with a glassy carbon electrode modified with microcrystals of the PmFT polymer conducted in aqueous (with KCl supporting electrolyte) or neat ionic liquid ([bmpyr][TFSA]) media also are reported. Different mechanisms apply in each of these cases.  相似文献   

3.
Structures of New Bis(pentafluorophenyl)halogeno Mercurates [{Hg(C6F5)2}3(μ‐X)] (X = Cl, Br, I) From the reactions of [PNP]Cl or [PPh4]Y (Y = Br, I) with Hg(C6F5)2 crystals of the composition [Cat][{Hg(C6F5)2}3X] (Cat = PNP, X = Cl ( 1 ); Cat = PPh4, X = Br ( 2 ), I ( 3 )) are formed. 1 crystallizes in the triclinic space group P1¯, 2 and 3 crystallize isotypically in the monoclinic space group C2/c. In the crystals the halide anions are surrounded by three Hg(C6F5)2 molecules. The reaction of [PPh4]Br with Hg(C6F5)2 under slightly changed conditions gives the compound [PPh4]2[{Hg(C6F5)2}3(μ‐Br)][{Hg(C6F5)2}2(μ‐Br)] ( 4 ).  相似文献   

4.
The reaction of Cp2Zr(OPri)2 with [H(OEt2)2][H2N{B(C6F5)3}2] in dichloromethane at room temperature gives [Cp2Zr(OPri)(HOPri)]+[H2N{B(C6F5)3}2] · Et2O in high yield. The crystal structure is reported. The complex contains a short Zr-alkoxide and a longer Zr-alcohol bond; the OH group of the coordinated isopropanol is hydrogen-bonded to a diethyl ether molecule. The complex initiates the polymerisation of propylene oxide, most probably via a cationic mechanism.  相似文献   

5.
[(BDI)Mg+][B(C6F5)4] ( 1 ; BDI=CH[C(CH3)NDipp]2; Dipp=2,6-diisopropylphenyl) was prepared by reaction of (BDI)MgnPr with [Ph3C+][B(C6F5)4]. Addition of 3-hexyne gave [(BDI)Mg+ ⋅ (EtC≡CEt)][B(C6F5)4]. Single-crystal X-ray analysis, NMR investigations, Raman spectra, and DFT calculations indicate a significant Mg-alkyne interaction. Addition of the terminal alkynes PhC≡CH or Me3SiC≡CH led to alkyne deprotonation by the BDI ligand to give [(BDI-H)Mg+(C≡CPh)]2 ⋅ 2 [B(C6F5)4] ( 2 , 70 %) and [(BDI-H)Mg+(C≡CSiMe3)]2 ⋅ 2 [B(C6F5)4] ( 3 , 63 %). Addition of internal alkynes PhC≡CPh or PhC≡CMe led to [4+2] cycloadditions with the BDI ligand to give {Mg+C(Ph)=C(Ph)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 4 , 53 %) and {Mg+C(Ph)=C(Me)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 5 , 73 %), in which the Mg center is N,N,C-chelated. The (BDI)Mg+ cation can be viewed as an intramolecular frustrated Lewis pair (FLP) with a Lewis acidic site (Mg) and a Lewis (or Brønsted) basic site (BDI). Reaction of [(BDI)Mg+][B(C6F5)4] ( 1 ) with a range of phosphines varying in bulk and donor strength generated [(BDI)Mg+ ⋅ PPh3][B(C6F5)4] ( 6 ), [(BDI)Mg+ ⋅ PCy3][B(C6F5)4] ( 7 ), and [(BDI)Mg+ ⋅ PtBu3][B(C6F5)4] ( 8 ). The bulkier phosphine PMes3 (Mes=mesityl) did not show any interaction. Combinations of [(BDI)Mg+][B(C6F5)4] and phosphines did not result in addition to the triple bond in 3-hexyne, but during the screening process it was discovered that the cationic magnesium complex catalyzes the hydrophosphination of PhC≡CH with HPPh2, for which an FLP-type mechanism is tentatively proposed.  相似文献   

6.
As the dysprosocenium complex [Dy(Cpttt)2][B(C6F5)4] (Cpttt=C5H2tBu3-1,2,4, 1-Dy ) exhibits magnetic hysteresis at 60 K, similar lanthanide (Ln) complexes have been targeted to provide insights into this remarkable property. We recently reported homologous [Ln(Cpttt)2][B(C6F5)4] ( 1-Ln ) for all the heavier Ln from Gd–Lu; herein, we extend this motif to the early Ln. We find, for the largest LnIII cations, that contact ion pairs [Ln(Cpttt)2{(C6F51-F)B(C6F5)3}] ( 1-Ln ; La–Nd) are isolated from reactions of parent [Ln(Cpttt)2(Cl)] ( 2-Ln ) with [H(SiEt3)2][B(C6F5)4], where the anion binds weakly to the equatorial sites of [Ln(Cpttt)2]+ through a single fluorine atom in the solid state. For smaller SmIII, [Sm(Cpttt)2][B(C6F5)4] ( 1-Sm ) is isolated, which like heavier 1-Ln does not exhibit equatorial anion interactions, but the EuIII analogue 1-Eu could not be synthesised due to the facile reduction of EuIII precursors to EuII products. Thus with the exception of Eu and radioactive Pm this work constitutes a structurally similar family of Ln metallocenium complexes, over 50 years after the [M(Cp)2]+ series was isolated for the 3d metals.  相似文献   

7.
Reduction of the R2P-functionalized zirconocene dichlorides [C5Me4(CH2)2PR2] (C5Me5)ZrCl2 (R = Me (1) and Ph (2)) and [C5Me4(CH2)2PMe2][C5Me4(CH2)2PR2]ZrCl2 (R = Me (3) and Ph (4)) with amalgamated magnesium was studied. In the reduction of compounds 1 and 2, intramolecular C-H activation highly selectively afforded the fulvene hydride complexes Zr(H)(η5−C5Me5)[η52(C,P)−(CH2)C5Me3CH2CH2PR2] (R = Me (7), Ph (8)); in the case of compound 2, the aryl hydride Zr(H)(η5:C5Me5)[η51(C)−C5Me4CH2CH2PPh(o−C6H4)] (9) was also formed. The reduction of complexes 3 and 4 gave the ZrII derivatives Zr[η51(P)− C5Me4CH2CH2PMe2]2 (12) and Zr[η51(P)−C5Me4CH2CH2PMe2][η51(P)−C5Me4CH2 CH2PPh2] (14) stabilized by two phosphine groups. The second product in the reduction of compound 4 was the fulvene hydride complex Zr(H)(η5−C5Me4CH2CH2PPh2)[η52(C,P)−(CH2)C5Me3CH2CH2PMe2] (15). The reaction of compound 7 with an excess of MeI resulted selectively in replacement of the hydride ligand by iodide to give the complex ZrI(η5−C5Me5)[η52(C,P)−(CH2)C5Me3CH2CH2PMe2] (10). In contrast, in the reaction of compound 7 with Me2Si(H)Cl, the Zr-CH2 bond underwent cleavage to give the chloride hydride complex Zr(H)Cl(η5−C5Me5)[η51(P)−C5Me3(CH2SiMe2H)CH2CH2PMe2] (11). In the reaction of complex 12 with CO, a phosphine group was replaced by CO to form the complex Zr(CO)(η5−C5Me4CH2CH2PMe2)[η51(P)−C5Me4CH2CH2PMe2] (13). The results obtained were compared with analogous reduction reactions of MeO-, MeS-, and Me2N-functionalized zirconocene dichlorides. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 65–74, January, 2008.  相似文献   

8.
Blue [{Cu(2,2′-bipy)2}2{α-SiW12O40}] (bipy = bipyridyl) (1) and pale yellow [Mn(2,2′-bipy)3]2[α-SiW12O40] (2) have been synthesized hydrothermally and characterized by IR spectroscopy and single crystal X-ray structure analysis. In 1, the [α-SiW12O40]4− ion acts as a bridge between the two [{Cu(2,2′-bipy)2]2+ moieties via coordination through the terminal oxygen atoms, while in 2, the [Mn(2,2′-bipy)3]2+ ion balances the charge on the polyoxo anion without forming any covalent bond. To the best of our knowledge, this is the first example of transition metal-mediated transformation of [α-SiW9O34]10− to [α-SiW12O40]4−. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

9.
A new complex compound bis[(dibenzo-18-crown-6)potassium]bis(μ2-chloro)-tetrachlorodicuprate( II), {[K(Db18c6)]2Cu2Cl6} (I) was prepared and its crystal structure was investigated by XRD analysis. Complex molecule I consist of anion [Cu2Cl6]2− located in a crystallographic center of inversion, and two centrosymmetrical to each other complex cations [K(Db18c6)]+ of “guest-host” type: the cation K+ is located in the cavity of the crown-ligand Db18c6 and is coordinated by all its six O atoms, and also by one Cl atom of anion [Cu2Cl6]2−. The coordination of this cation K+ is enlarged up to hexagonal-bipyramidal due to the formation of unusual coordination bond K+ → π(   相似文献   

10.
1H NMR method showed that in systems based on triisobutylaluminum (TIBA) and triphenylcyclopropenylium [Ph3C3]+[B(C6F5)4]–(CPB) or triphenylmethylium [Ph3C]+[B(C6F5)4]–(TB) perfluorophenylborates in a toluene–dichloromethane mixture the Friedel–Crafts process occurs with the formation of ditolylmethane (DTM) accompanied by the complete decomposition of TIBA to form isobutane. 19F NMR spectroscopy showed that the [B(C6F5)4]–anion decomposes in the systems to form B(C6F5)3 and HC6F5. The short-living [AlBu2 i]+ cation formed in the reaction of perfluorophenylborates with TIBA is assumed to be the species initiating the process. It has been shown that CPB is less reactive than TB. The addition of a stoichiometric amount of Ph2CCpFluHfMe2 exerts no effect on the process with the CPB-containing system but inhibits the reaction in the case of TB.  相似文献   

11.
Room‐temperature ionic liquids containing the triply charged tricyanomelaminate (tcmel) ion [C3N6(CN)3]3? were synthesized. The 1‐methyl‐3‐methylimidazolium (MMIm), 1‐ethyl‐3‐methylimidazolium (EMIm), and 1‐butyl‐3‐methylimidazolium (BMIm) salts of the tricyanomelaminate ion have glass transition temperatures (?6, ?20, and ?30 °C) similar to those found for the analogous monomeric dicyanoamide salts. They are thermally stable up to over 200 °C and dissolve in polar organic solvents. Addition of B(C6F5)3 to M3[tcmel] (M=Na, MMIm, EMIm, BMIm) yields salts containing the very voluminous adduct ion [C3N6{CN ? B(C6F5)3}3]3? (tcmel_3B). The solid‐state structure of [MMIm]3[tcmel] shows only long cation ??? anion contacts but in large number, while the solid‐state structure of [Na(THF)3]3[tcmel_3B] ? 1.76 THF displays strong interactions of the sodium cation with the amido nitrogen atoms of the anion. Hence this adduct anion cannot be regarded as a weakly coordinating anion. A similar situation is found for the MMIm salt, [MMIm]3[tcmel_3B] ? 2.66 CH2Cl2, in which weak hydrogen bonds with the acidic proton of the MMIm ion are observed. On the basis of computations the energetics, structural trends, and charge transfer of adduct anion formation were studied.  相似文献   

12.
New cyclosiloxanolate transition metal cluster complex derivatives were prepared. PhSiO2K reacted with NiX2 (X2 = Cl2 or acac) to give K2{[η6−(PhSiO2)6]23−(OH)]2Ni4K4}, a mixed group 1–group 10 metal complex. PhSiO2Na reacted with Ni(NH3)6I2 to give Na{[η6−(PhSiO2)6]2Ni66−I)} as the first example of “encapsulated” I ion in siloxanolate complexes. The macrocyclic Na4{[η12−(PhSiO2)12]Cu4} complex reacted with η6−(1,3,5−C7H8)Cr(CO)3 to give the heterobimetallic adduct Na4{[η12−(PhSiO2)12]Cu4}· [Cr(CO)3]3 as one of the rare examples of heterobimetallic complexes with different oxidation numbers of the metals. The copper derivative {[η6−(PhSiO2)6]2Cu6(n−BuOH)5} reacted in MeOH/CHCl3 (1:6) with Et4NCN to give the hexanuclear complex {[η6−(PhSiO2)6]2Cu62−C3H5N2O2)2}, containing 2-amino-2-oxoetanimidic acid methyl ester monoanion ligands, product of an unexpected C–C coupling reaction. This latter complex was characterized also by X-ray diffraction crystal and molecular structure determination. This paper is dedicated to the 70th birthday of Professor Dr. Gunter Schmid (Essen), pioneer of large cluster chemistry, known to friends as GOLD-Schmid, because of his famous discovery of the Au55 cluster. The Authors are proud to be within his many friends.  相似文献   

13.
The reaction of [Pt2(μ-S)2(P-P)2] (P-P=2PPh3, 2PMe2Ph, dppf) [dppf=1,1-bis(diphenylphosphino)ferrocene] with cis-[M(C6F5)2(PhCN)2] (M=Ni, Pd) or cis-[Pt(C6F5)2(THF)2] (THF=tetrahydrofuran) afforded sulfide aggregates of the type [{Pt23-S)2(P-P)2}M(C6F5)2] (M=Ni, Pd, Pt). X-ray crystal analysis revealed that [{Pt23-S)2(dppf)2}Pd(C6F5)2], [{Pt23-S)2(PPh3)2}Ni(C6F5)2], [{Pt23-S)2(PPh3)2}Pd(C6F5)2] and [{Pt23-S)2(PMe2Ph)2}Pt(C6F5)2] have triangular M3S2 core structures capped on both sides by μ3-sulfido ligands. The structural features of these polymetallic complexes are described. Some of them display short metal-metal contacts.  相似文献   

14.
Five novel 2,3-naphtho crown ether group 10 metal bis(dithiolate) complexes, [Na(N15C5)2]2[Pd(mnt)2] (1), [Na(N15C5)]2[Pd(i-mnt)2] (2) and [K(N18C6)]2[M(i-mnt)2] (3 5) (where mnt = 1,2-dicyanoethylene-1,2-dithiolate, i-mnt = 1,1-dicyanoethylene-2,2-dithiolate and M = Ni, Pd, Pt for complexes 35, respectively), have been synthesized and characterized by elemental analysis, FT-IR, UV–Visible spectra and single crystal X-ray diffraction. X-ray diffraction analyses reveal that complexes 1 and 2 have different structural features while complexes 35 are structurally isomorphous. Complex 1 consists of two [Na(N15C5)2]+ sandwich complex cations and one [Pd(mnt)2]2− anion, affording a zero-dimensional structure. For 2, the [Na(N15C5)]+ mono-capped complex cations act as the bridges linking the [Pd(i-mnt)2]2− anions into a 1D infinite chain through Na–N interactions and SȮFC and SȮFπ interactions are observed in the resulting chain. Complexes 35 all consist of two [K(N18C6)]+ complex cations and one [M(i-mnt)2]2− (M = Ni, Pd or Pt) anion and the complex molecules are linked into␣1D␣chains by the bridging K–O(ether) interactions between the adjacent [K(N18C6)]+ units. What’s novel is that the resulting chains are assembled into novel 2D networks through interchain π–π stacking interactions between the neighboring naphthylene moieties of N18C6. The stack model of naphthylene group in complexes 35 is discussed.  相似文献   

15.
The new heteroleptic mercury(II) complex PhHgN(SiMe3)2(1) reacts with the strong Brønsted acid [H(OEt2)2][H2N{B(C6F5)3}2] with cleavage of a N-Si bond to give [C6H5Hg(H2NSiMe3)][H2N{B(C6F5)3}2] (2), a phenyl-mercury(II) cation stabilised by a primary amine and a non-coordinating counter-anion. Attempts to generate donor-free aryl mercury cations were not successful. The crystal structure of 2 · CH2Cl2 shows short π-bonding interactions between the metal and the phenyl ring of a neighbouring cation; the geometry about the mercury(II) atom is nearly linear. The X-ray structures of the new salts [H2N(SiMe3)2 · H3NSiMe3][B(C6F5)4]2 and [Et3O][H2N{B(C6F5)3}2] · CH2Cl2 are also presented.  相似文献   

16.
Molybdenum(VI) bis(imido) complexes [Mo(NtBu)2(LR)2] (R=H 1 a ; R=CF3 1 b ) combined with B(C6F5)3 ( 1 a /B(C6F5)3, 1 b /B(C6F5)3) exhibit a frustrated Lewis pair (FLP) character that can heterolytically split H−H, Si−H and O−H bonds. Cleavage of H2 and Et3SiH affords ion pairs [Mo(NtBu)(NHtBu)(LR)2][HB(C6F5)3] (R=H 2 a ; R=CF3 2 b ) composed of a Mo(VI) amido imido cation and a hydridoborate anion, while reaction with H2O leads to [Mo(NtBu)(NHtBu)(LR)2][(HO)B(C6F5)3] (R=H 3 a ; R=CF3 3 b ). Ion pairs 2 a and 2 b are catalysts for the hydrosilylation of aldehydes with triethylsilane, with 2 b being more active than 2 a . Mechanistic elucidation revealed insertion of the aldehyde into the B−H bond of [HB(C6F5)3]. We were able to isolate and fully characterize, including by single-crystal X-ray diffraction analysis, the inserted products Mo(NtBu)(NHtBu)(LR)2][{PhCH2O}B(C6F5)3] (R=H 4 a ; R=CF3 4 b ). Catalysis occurs at [HB(C6F5)3] while [Mo(NtBu)(NHtBu)(LR)2]+ (R=H or CF3) act as the cationic counterions. However, the striking difference in reactivity gives ample evidence that molybdenum cations behave as weakly coordinating cations (WCC).  相似文献   

17.
The alkoxido-titanium pentamolybdate [(iPrO)TiMo5O18]3− (1) has been obtained as its tetrabutylammonium (TBA) salt by hydrolysis of a mixture containing (TBA)2[Mo2O7], (TBA)4[Mo8O26] and Ti(OiPr)4 in MeCN and has been characterised by 1H, 13C, 17O, 49Ti and 95Mo NMR and FTIR spectroscopy, electrospray ionisation mass spectrometry, elemental microanalysis and single-crystal X-ray crystallography. The Lindqvist-type structure is derived from [Mo6O19]2− by replacement of {Mo=O}4+ by {(iPrO)Ti}3+ and shows bond alternation in the TiMo3O4 rings, with average bond distances of 1.956(8) ? for Ti–O(Mo), 1.832(7) ? for Mo–O(Ti), 1.943(7) ? for Moeq–O(Moax) and 1.910(6) ? for Moax–O(Moeq), while the increase in charge results in a decrease in 17O NMR chemical shift for terminal Mo=O groups from δ 933 for [Mo6O19]2− to δ 875 and 857 for 1 and a shift in νMo=O from 951 cm−1 for [Mo6O19]2− to 930 cm−1 for 1. The main peaks in the negative-ion electrospray ionisation mass spectrum of (TBA)3 1 could be assigned to ion aggregates containing 1 or fragments derived from 1, including {(TBA)2[(iPrO)TiMo5O18]}, {(TBA)[(iPrO)TiMo5O18]}2−, {(iPrO)TiMo2O8}, {TiMo5O18}2−, {TiMo4O15}2− and {Mo3O10}2−.  相似文献   

18.
The synthesis and full characterization of α-silylated (α-SiCPs; 1 – 7 ) and α-germylated (α-GeCPs; 11 – 13 ) phosphorus ylides bearing one chloride substituent R3PC(R1)E(Cl)R22 (R=Ph; R1=Me, Et, Ph; R2=Me, Et, iPr, Mes; E=Si, Ge) is presented. The molecular structures were determined by X-ray diffraction studies. The title compounds were applied in halide abstraction studies in order to access cationic species. The reaction of Ph3PC(Me)Si(Cl)Me2 ( 1 ) with Na[B(C6F5)4] furnished the dimeric phosphonium-like dication [Ph3PC(Me)SiMe2]2[B(C6F5)4]2 ( 8 ). The highly reactive, mesityl- or iPr-substituted cationic species [Ph3PC(Me)SiMes2][B(C6F5)4] ( 9 ) and [Ph3PC(Et)SiiPr2][B(C6F5)4] ( 10 ) could be characterized by NMR spectroscopy. Carrying out the halide abstraction reaction in the sterically demanding ether iPr2O afforded the protonated α-SiCP [Ph3PCH(Et)Si(Cl)iPr2][B(C6F5)4] ( 6 dec ) by sodium-mediated basic ether decomposition, whereas successfully synthesized [Ph3PC(Et)SiiPr2][B(C6F5)4] ( 10 ) readily cleaves the F−C bond in fluorobenzene. Thus, the ambiphilic character of α-SiCPs is clearly demonstrated. The less reactive germanium analogue [Ph3PC(Me)GeMes2][B{3,5-(CF3)2C6H3}4] ( 14 ) was obtained by treating 11 with Na[B{3,5-(CF3)2C6H3}4] and fully characterized including by X-ray diffraction analysis. Structural parameters indicate a strong CYlide−Ge interaction with high double bond character, and consequently the C−E (E=Si, Ge) bonds in 9 , 10 and 14 were analyzed with NBO and AIM methods.  相似文献   

19.
The reactivity of aryl‐substituted stannylenes, Ar2Sn ( 4 ), towards silylarenium borates, [R3SiArH][B(C6F5)4] ( 3 ), was investigated. The reaction with 2,3,4‐trimethyl‐6‐tert‐butylphenyl (mebp)‐substituted stannylene gave silyl‐substituted stannylium ions 2 a , b , which were characterized by NMR spectroscopy supported by the results of quantum‐mechanical computations of molecular structures and magnetic properties. The tri‐iso‐propylphenyl‐substituted stannylium ions 2 c , d undergo a decomposition reaction in toluene to give the dicationic tin–arene complex [Sn(C7H8)3]2+ ( 5 ) in the form of the [B(C6F5)4] salt in high yields. The 5 [B(C6F5)4]2 salt was identified by single crystal X‐ray diffraction analysis and by Mössbauer spectroscopy. The bonding situation was investigated by using natural bond orbital (NBO) and quantum theory of atoms in molecules (QTAIM) calculations. The substitution of the weakly coordinating borate anion by the carboranate [CB11H6Br6]? results in replacement of the toluene ligands and formation of tin(II) carboranate with only weak Sn2+–anion interactions as suggested by the solid‐state structure of the isolated salt.  相似文献   

20.
The benzene complex [1-(η-C6H6)-12-ButNH-1,2,4,12-FeC3B8H10]+ (3a) was synthesized by the photochemical reaction of [(η5-C6H7)Fe(η-C6H6)]+ (1) with the anion [7-ButNH-7,8,9-C3B8H10] followed by the treatment of ferracarborane 1-(η5-C6H7)-12-ButNH-1,2,4,12-FeC3B8H10 (2) with hydrochloric acid. The benzene ligand in cation 3a is replaced by alkyl-substituted benzenes under visible light irradiation in CH2Cl2 to form [1-(η-C6R6)-12-ButNH-1,2,4,12-FeC3B8H10]+ (3b–e; C6R6 is toluene (b), mesitylene (c), hexamethylbenzene (d), or anisole (e)). The structure of [3c]PF6 was established by X-ray diffraction.  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