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.
Kinetically stabilized congeners of carbenes, R2C, possessing six valence electrons (four bonding electrons and two non‐bonding electrons) have been restricted to Group 14 elements, R2E (E=Si, Ge, Sn, Pb; R=alkyl or aryl) whereas isoelectronic Group 15 cations, divalent species of type [R2E]+ (E=P, As, Sb, Bi; R=alkyl or aryl), were unknown. Herein, we report the first two examples, namely the bismuthenium ion [(2,6‐Mes2C6H3)2Bi][BArF4] ( 1 ; Mes=2,4,6‐Me3C6H2, ArF=3,5‐(CF3)2C6H3) and the stibenium ion [(2,6‐Mes2C6H3)2Sb][B(C6F5)4] ( 2 ), which were obtained by using a combination of bulky meta‐terphenyl substituents and weakly coordinating anions.  相似文献   

2.
A metal‐containing N‐heterocyclic germylene based on a N‐mesityl (Mes)‐substituted oxalamidine framework is reported. The precursor (MesN=)2C–C(–N(H)Mes)2 ( 1 H2) was converted into its rhodium complex [Rh(κ2N‐ 1 H2)(cod)][OTf] ( 2 ) (cod = 1,5‐cyclooctadiene; OTf = triflate) in 62 % isolated yield. Subsequent reaction of 2 with Ge{N(SiMe3)2}2 gave the crystalline N‐heterocyclic germylene [Rh(cod)(μ‐ 1 )Ge][OTf] ( 3 ) in 50 % yield. The compounds under study were fully characterized by various methods, also including X‐ray crystallographic studies on single crystals of 2 and 3 . Density functional theory (DFT) calculations revealed that π conjugation in the bridging oxalamidine framework is increased and n(N)–p(Ge) π bonding is decreased upon κ2N metal coordination; a further weakening of the Ge–N bond occurs through triflate coordination to the GeII atom. Nevertheless, preliminary coordination studies revealed that 3 behaves as 2‐electron (L ‐type) germylene donor ligand. Treatment of 3 with [Ir(cod)Cl]2 furnished the heterobimetallic complex [Rh(cod)(μ‐ 1 )Ge‐Ir(cod)Cl][OTf] ( 4 ), as evidenced by NMR spectroscopic investigations and DFT calculations.  相似文献   

3.
Transition metal tetrylene complexes offer great opportunities for molecular cooperation due to the ambiphilic character of the group 14 element. Here we focus on the coordination of germylene [(ArMes2)2Ge :] (ArMes=C6H3-2,6-(C6H2-2,4,6-Me3)2) to [RhCl(COD)]2 (COD=1,5-cyclooctadiene), which yields a neutral germyl complex in which the rhodium center exhibits both η6- and η2-coordination to two mesityl rings in an unusual pincer-type structure. Chloride abstraction from this species triggers a singular dehydrogenative double C−H bond activation across the Ge/Rh motif. We have isolated and fully characterized three rhodium-germyl species associated to three C−H cleavage events along this process. The reaction mechanism has been further investigated by computational means, supporting the key cooperative action of rhodium and germanium centers.  相似文献   

4.
Reaction of a lithium boryl, [(THF)2Li{B(DAB)}] (DAB=[(DipNCH)2]2?, Dip=2,6‐diisopropylphenyl), with a dinuclear magnesium(I) compound [{(MesNacnac)Mg}2] (MesNacnac=[HC(MeCNMes)2]?, Mes=mesityl) unexpectedly afforded a rare example of a terminal magnesium boryl species, [(MesNacnac)(THF)Mg{B(DAB)}]. Attempts to prepare the magnesium boryl via a salt metathesis reaction between the lithium boryl and a β‐diketiminato magnesium iodide compound, instead led to an intractable mixture of products. Similarly, reaction of the lithium boryl with a β‐diketiminato beryllium bromide precursor, [(DepNacnac)BeBr] (Dep=2,6‐diethylphenyl) did not give a beryllium boryl, but instead afforded an unprecedented example of a beryllium substituted diazaborole heterocycle, [{(DepNacnac)Be(4‐DAB?H)}BBr]. For sake of comparison, the same group 2 halide precursor compounds were treated with a potassium gallyl analogue of the lithium boryl, viz. [(tmeda)K{:Ga(DAB)}] (tmeda=N,N,N’,N’‐tetramethylethylenediamine), but no reactions were observed.  相似文献   

5.
Addition of MesN3 (Mes=2,4,6-Me3C6H2) to germylene [(NONtBu)Ge] (NONtBu=O(SiMe2NtBu)2) ( 1 ) gives germanimine, [(NONtBu)Ge=NMes] ( 2 ). Compound 2 behaves as a metalloid, showing reactivity reminiscent of both transition metal-imido complexes, undergoing [2+2] addition with heterocumulenes and protic sources, as well as an activated diene, undergoing a [4+2] cycloaddition, or “metallo”-Diels–Alder, reaction. In the latter case, the diene includes the Ge=N bond and π-system of the Mes substituent, which is reactive towards dienophiles including benzaldehyde, benzophenone, styrene, and phenylacetylene.  相似文献   

6.
Reduction of a variety of extremely bulky amido Group 12 metal halide complexes, [LMX(THF)0,1] (L=amide; M=Zn, Cd, or Hg; X=halide) with a magnesium(I) dimer gave a homologous series of two‐coordinate metal(I) dimers, [L′MML′] (L′=N(Ar?)(SiMe3), Ar?=C6H2{C(H)Ph2}2Pri‐2,6,4); and the formally zinc(0) complex, [L*ZnMg(MesNacnac)] (L*=N(Ar*)(SiPri3); Ar*=C6H2{C(H)Ph2}2Me‐2,6,4; MesNacnac=[(MesNCMe)2CH]?, Mes=mesityl), which contains the first unsupported Zn? Mg bond. Two equivalents of [L*ZnMg(MesNacnac)] react with ZnBr2 or ZnBr2(tmeda) to give the mixed valence, two‐coordinate, linear tri‐zinc complex, [L*ZnIZn0ZnIL*], and the first zinc(I) halide complex, [L*ZnZnBr(tmeda)], respectively. The analogues [L*ZnMZnL*] (M=Cd or Hg), were also prepared, the Cd species contains the first Zn? Cd bond in a molecular compound. Metal–metal bonding was studied by DFT calculations.  相似文献   

7.
Reaction of a labile tungsten nitrile complex, [(Cp*)W(CO)2(NCMe)Me] (Cp*=η5‐C5Me5), with H3SiC(SiMe3)3 gave the hydrido(hydrosilylene) complex [(Cp*)(CO)2(H)W?Si(H){C(SiMe3)3}] ( 1a ). The hydrido(silylene) complex [(η5‐C5Me4Et)(CO)2(H)W?SiMes2] ( 2 ) (Mes=2,4,6‐Me‐C6H2) was synthesized by a similar reaction with H2SiMes2. There is a strong interligand interaction between the hydrido and silylene ligands of these complexes; this was confirmed by a neutron diffraction study of [D2] 1b , that is, the deuterido and η5‐C5Me4Et derivative of 1a . The exchange between the W? H and the Si? D groups was observed in the deuterido complex [D] 1a . This H/D exchange proceeded slowly at room temperature, but very rapidly under UV irradiation. Variable‐temperature NMR spectroscopy measurements show the dynamic behavior of carbonyl ligands in 1a . Complex 1a reacted with acetone at room temperature to give mainly a hydrosilylation product, [(Cp*)(CO)2(H)W?Si(OiPr){C(SiMe3)3}] ( 3a ), along with a siloxy complex, [(Cp*)(CO)2WO(Si(H)iPr{C(SiMe3)3})] ( 4a ). At low temperature, a different reaction, namely, α‐H abstraction, proceeded to give an equilibrium mixture of 1a and a dihydrido(silyl) complex, [(Cp*)(CO)2(H)2W(Si(H){OC(?CH2)Me}{C(SiMe3)3})] ( 5 ).  相似文献   

8.
Boragermene 3 featuring a double bond between the Ge and dicoordinate B atoms has been synthesized for the first time by reacting the cyclic (alkyl)(boryl)germylene–PMe3 adduct 1 with Cl2BN(SiMe3)2 followed by reductive dehalogenation with KC8. Addition of a Lewis base (MeNHC) to 3 leads to the formation of the corresponding adduct 4 , which shows double bond character between the Ge and tricoordinate B atoms. Compound 3 undergoes hydrogenation with H2 concomitant with a complete scission of the Ge=B bond.  相似文献   

9.
Boragermene 3 featuring a double bond between the Ge and dicoordinate B atoms has been synthesized for the first time by reacting the cyclic (alkyl)(boryl)germylene–PMe3 adduct 1 with Cl2BN(SiMe3)2 followed by reductive dehalogenation with KC8. Addition of a Lewis base (MeNHC) to 3 leads to the formation of the corresponding adduct 4 , which shows double bond character between the Ge and tricoordinate B atoms. Compound 3 undergoes hydrogenation with H2 concomitant with a complete scission of the Ge=B bond.  相似文献   

10.
Despite the explosive growth of germylene compounds as ligands in transition metal complexes, there is a modicum of precedence for the germylene zinc complexes. In this work, the synthesis and characterization of new germylene zinc complexes [PhC(NtBu)2Ge{N(SiMe3)2}→ZnX2]2 (X= Br ( 2 ) and I ( 3 )) supported by (benz)‐amidinato germylene ligands are reported. The solid‐state structures of 2 and 3 have been validated by single‐crystal X‐ray diffraction studies, which revealed the dimeric nature of the complexes, with distorted tetrahedral geometries around the Ge and Zn center. DFT calculations reveal that the Ge–Zn bonds in 2 and 3 are dative in nature. The reaction of 2 with elemental sulfur resulted in the first structurally characterized germathione stabilized ZnBr2 complexes PhC(NtBu)2Ge(=S){N(SiMe3)2}→ZnBr2 ( 5 ). Therefore, the Ge=S in 5 is in‐between Ge–S single and Ge=S double bond length, owing to the coordination of a sulfur lone pair of electrons to ZnBr2.  相似文献   

11.
Treatment of dichloromethyl‐tris(trimethylsilyl)silane (Me3Si)3Si–CHCl2 ( 1 ), prepared by the reaction of tris(trimethylsilyl)silane with chloroform in presence of potassium tertbutoxide, with organolithium reagents (molar ratio 1 : 3) affords the bis(trimethylsilyl)methyl‐disilanes Me3SiSiR2–CH(SiMe3)2 ( 12 a–d ) ( a : R = Me, b : R = n‐Bu, c : R = Ph, d : R = Mes). The formation of 12 a–d is discussed as proceeding through an exceptional series of isomerization and addition reactions involving intermediate silyl substituted carbenoids and transient silenes. The carbenoid (Me3Si)2PhSi–C(SiMe3)LiCl ( 8 c ) is moderately stable at low temperature and was trapped with water to give (Me3Si)2PhSi–CH(SiMe3)Cl ( 9 c ) and with chlorotrimethylsilane affording (Me3Si)2PhSi–CCl(SiMe3)2 ( 7 c ). For 12 d an X‐ray crystal structure analysis was performed, which characterizes the compound as a highly congested silane with bond parameters significantly deviating from standard values.  相似文献   

12.
The synthesis of a boryl-substituted germanium(II) cation, [Ge{B(NDippCH)2}(IPrMe)]+, (Dipp=2,6-diisopropylphenyl) featuring a supporting N-heterocyclic carbene (NHC) donor, has been explored through chloride abstraction from the corresponding (boryl)(NHC)GeCl precursor. Crystallographic studies in the solid state and UV/Vis spectra in fluorobenzene solution show that this species dimerizes under such conditions to give [(IPrMe){(HCNDipp)2B}Ge=Ge{B(NDippCH)2}(IPrMe)]2+ (IPrMe = 1,3-diisopropyl-4,5-dimethylimidazolin-2-ylidene), which can be viewed as an imidazolium-functionalized digermene. The dimer is cleaved in the presence of donor solvents such as [D8]thf or [D5]pyridine, to give monomeric adducts of the type [Ge{B(NDippCH)2}(IPrMe)(L)]+. In the case of the thf adduct, the additional donor is shown to be sufficiently labile that it can act as a convenient in situ source of the monomeric complex [Ge{B(NDippCH)2}(IPrMe)]+ for oxidative bond-activation chemistry. Thus, [Ge{B(NDippCH)2}(IPrMe)(thf)]+ reacts with silanes and dihydrogen, leading to the formation of GeIV products, whereas the cleavage of the N−H bond in ammonia ultimately yields products containing C−H and B−N bonds. The facile reactivity observed in E−H bond activation is in line with the very small calculated HOMO–LUMO gap (132 kJ mol−1).  相似文献   

13.
An experimental and theoretical study of the first compound featuring a Si?P bond to a two‐coordinate silicon atom is reported. The NHC‐stabilized phosphasilenylidene (IDipp)Si?PMes* (IDipp=1,3‐bis(2,6‐diisopropylphenyl)imidazolin‐2‐ylidene, Mes*=2,4,6‐tBu3C6H2) was prepared by SiMe3Cl elimination from SiCl2(IDipp) and LiP(Mes*)SiMe3 and characterized by X‐ray crystallography, NMR spectroscopy, cyclic voltammetry, and UV/Vis spectroscopy. It has a planar trans‐bent geometry with a short Si? P distance of 2.1188(7) Å and acute bonding angles at Si (96.90(6)°) and P (95.38(6)°). The bonding parameters indicate the presence of a Si?P bond with a lone electron pair of high s‐character at Si and P, in agreement with natural bond orbital (NBO) analysis. Comparative cyclic voltammetric and UV/Vis spectroscopic experiments of this compound, the disilicon(0) compound (IDipp)Si?Si(IDipp), and the diphosphene Mes*P?PMes* reveal, in combination with quantum chemical calculations, the isolobal relationship of the three double‐bond systems.  相似文献   

14.
Aryldimethylsilyl‐substituted hypersilanes ArSiMe2Hyp [Hyp =Si(SiMe3)3; Ar = Mes ( 3a ), Bph ( 3b ), Mph ( 3c ), Pph ( 3d ), Tph ( 3e ); Dpp ( 3f ); with Mes = 2,4,6‐Me3C6H2, Bph = 2‐PhC6H4, Mph = 2‐MesC6H4, Pph = 2′,3′,4′,5′,6′‐pentamethylbiphenyl2‐yl, Tph = 2′,4′,6′‐triisopropylbiphenyl‐2‐yl and Dpp = 2,6‐Ph2C6H3] have been synthesized by a multi‐step reaction starting from the readily available starting materials ArI and (thf)xKHyp. For this purpose the aryl iodides were first converted into the lithium aryls by lithium‐halogen exchange with n‐butyl lithium. A salt‐metathesis reaction with SiMe2Cl2 gave the corresponding arylchlorodimethylsilanes ( 2a – e ) in excellent yields. Finally, reaction of 2a – e with (thf)xKHyp afforded the hypersilanes 3a – e in yields up to 86 %. Bis(aryldimethylsilyl)‐substituted silanes (ArSiMe2)2Si(SiMe3)2 (Ar = Mph ( 4c ), Tph ( 4e )) were obtained by the reaction of 3c or 3e with KOtBu and the corresponding chlorosilanes 2c or 2e . All new compounds have been characterized by melting point, elemental analysis, 1H, 13C, and 29Si NMR spectroscopy and for selected species by IR spectroscopy or mass spectrometry. In addition, the solid‐state structures of silanes 3a – d , 3f , and 4c have been investigated by single crystal X‐ray diffraction.  相似文献   

15.
The dinuclear palladium(I) complexes [L(Ar2HGe)Pd(μ‐GeAr2)2Pd(GeHAr2)L] (Ar=Ph, p‐Tol; L=PMe3, tBuNC) contain terminal germyl and bridging germylene ligands with the experimentally observed Ge???Ge bond lengths of 2.8263(4) Å (L=PMe3) and 2.928(1) Å (L=tBuNC), which are close to the longest Ge? Ge bond reported to date [2.714(1) Å]. Significant Ge???Ge interactions between the germylene and germyl ligands (PMe3 complexes > tBuNC complexes) are supported by DFT calculations, Wiberg bond indices (WBI), and natural bond orbital (NBO) analyses. Exchanging tBuNC for PMe3 ligands increases the Ge???Ge interaction, and simultaneously activates two Pd? Ge bonds. Adding the chelating diphosphine 1,2‐bis(diethylphosphino)ethane (depe) to the PMe3 complexes results in the intramolecular coupling of germyl and germylene ligands followed by extrusion of a digermane.  相似文献   

16.
Reactions of the beryllium dihalide complexes [BeX2(OEt2)2] (X=Br or I) with N,N,N′,N′‐tetramethylethylenediamine (TMEDA), a series of diazabutadienes, or bis(diphenylphosphino)methylene (DPPM) have yielded the chelated complexes, [BeX2(TMEDA)], [BeX2{(RN=CH)2}] (R=tBu, mesityl (Mes), 2,6‐diethylphenyl (Dep) or 2,6‐diisopropylphenyl (Dip)), and the non‐chelated system, [BeI21P‐DPPM)2]. Reactions of lithium or potassium salts of a variety of β‐diketiminates have given both three‐coordinate complexes, [{HC(RCNAr)2}BeX] (R=H or Me; Ar=Mes, Dep or Dip; X=Br or I); and four‐coordinate systems, [{HC(MeCNPh)2}BeBr(OEt2)] and [{HC(MeCNDip)(MeCNC2H4NMe2}BeI]. Alkali metal salts of ketiminate, guanidinate, boryl/phosphinosilyl amide, or terphenyl ligands, lead to dimeric [{BeI{μ‐[(OCMe)(DipNCMe)]CH}}2], and monomeric [{iPr2NC(NMes)2}BeI(OEt2)], [κ2N,P‐{(HCNDip)2B}(PPh2SiMe2)NBeI(OEt2)] and [{C6H3Ph2‐2,6}BeBr(OEt2)], respectively. Compound [{HC(MeCNPh)2}BeBr(OEt2)] undergoes a Schlenk redistribution reaction in solution, affording the homoleptic complex, [{HC(MeCNPh)2}2Be]. The majority of the prepared complexes have been characterized by X‐ray crystallography and multi‐nuclear NMR spectroscopy. The structures and stability of the complexes are discussed, as is their potential for use as precursors in poorly developed low oxidation state beryllium chemistry.  相似文献   

17.
Ge[N(SiMe2iPr)2]2: A New Germylene and its Coordination Chemistry leads to the Shortest Ge–Co Bond The high temperature reaction of Germanium with HBr, which is used for the synthesis of GeBr leads at a higher reaction pressure to a mixture of GeBr and GeBr2. GeBr2 reacts with the lithiumsalt LiN(SiMe2iPr)2 in good yield to the germylene Ge[N(SiMe2iPr)2]2. Subsequent reaction of this germylene with Co2(CO)8 leads to the germanium cobalt cluster compound {Co(CO)3Ge[N(SiMe2iPr)2]2}2 featuring the shortest Ge–Co bond giving hints to a possibly multiple bonded system.  相似文献   

18.
Tris(trimethylsilyl)silyllithium ( 3 ) reacted with aldehydes and ketones (molar ratio 2 : 1) according to a modified Peterson mechanism under formation of transient silenes, which were immediately trapped by excess 3 to give the organolithium derivatives (Me3Si)3SiSi(SiMe3)2C(Li)R1R2 ( 7 ). Hydrolysis of 7 afforded the alkylpolysilanes (Me3Si)3SiSi(SiMe3)2CHR1R2 ( 8 ). Depending on the substituents R1 and R2, 7 proved to be rather unstable in THF solution and underwent a rapid rearrangement, involving a 1,3‐Si,C‐trimethylsilyl migration, resulting in the formation of the lithium silanides (Me3Si)2Si(Li)Si(SiMe3)2C(SiMe3)R1R2 ( 9 ), which were hydrolized during the aqueous workup to give the H‐silanes (Me3Si)2Si(H)Si(SiMe3)2C(SiMe3)R1R2 ( 10 ). Reaction of 9 with chlorotrimethylsilane produced the 1‐trimethylsilylalkylpolysilanes (Me3Si)3SiSi(SiMe3)2C(SiMe3)R1R2 ( 11 ). The structures of the products described were elucidated by comprehensive spectral analyses. The results of X‐ray crystal structure analyses, performed for 8 l (R1 = H, R2 = 2,4,6‐(MeO)3C6H2), 10 d (R1 = H, R2 = Mes) and 11 d (R1 = H, R2 = Mes) are discussed and confirm the expected extreme sterical congestion of the molecules.  相似文献   

19.
The synthesis of an N‐heterocyclic silylene‐stabilized digermanium(0) complex is described. The reaction of the amidinate‐stabilized silicon(II) amide [LSiN(SiMe3)2] ( 1 ; L=PhC(NtBu)2) with GeCl2?dioxane in toluene afforded the SiII–GeII adduct [L{(Me3Si)2N}Si→GeCl2] ( 2 ). Reaction of the adduct with two equivalents of KC8 in toluene at room temperature afforded the N‐heterocyclic carbene silylene‐stabilized digermanium(0) complex [L{(Me3Si)2N}Si→ Ge?Ge←Si{N(SiMe3)2}L] ( 3 ). X‐ray crystallography and theoretical studies show conclusively that the N‐heterocyclic silylenes stabilize the singlet digermanium(0) moiety by a weak synergic donor–acceptor interaction.  相似文献   

20.
Quantum chemical calculations using density functional theory with the TPSS+D3(BJ) and M06‐2X+D3(ABC) functionals have been carried out to understand the mechanisms of catalyst‐free hydrogermylation/hydrostannylation reactions between the two‐coordinate hydrido‐tetrylenes :E(H)(L+) (E=Ge or Sn, L+=N(Ar+)(SiiPr3); Ar+=C6H2{C(H)Ph2}2iPr‐2,6,4) and a range of unactivated terminal (C2H3R, R=H, Ph, or tBu) and cyclic [(CH)2(CH2)2(CH2)n, n=1, 2, or 4] alkenes. The calculations suggest that the addition reactions of the germylenes and stannylenes to the cyclic and acyclic alkenes occur as one‐step processes through formal [2+2] addition of the E?H fragment across the C?C π bond. The reactions have moderate barriers and are weakly exergonic. The steric bulk of the tetrylene amido groups has little influence on the activation barriers and on the reaction energies of the anti‐Markovnikov pathway, but the Markovnikov addition is clearly disfavored by the size of the substituents. The addition of the tetrylenes to the cyclic alkenes is less exergonic than the addition to the terminal alkenes, which agrees with the experimentally observed reversibility of the former reactions. The hydrogermylation reactions have lower activation energies and are more exergonic than the stannylene addition. An energy decomposition analysis of the transition state for the hydrogermylation of cyclohexene shows that the reaction takes place with simultaneous formation of the Ge?C and (Ge)H?C′ bonds. The dominant orbitals of the germylene are the σ‐type lone pair MO of Ge, which serves as a donor orbital, and the vacant p(π) MO of Ge, which acts as acceptor orbital for the π* and π MOs of the olefin. Inspection of the transition states of some selected reactions suggests that the differences between the activation energies come from a delicate balance between the deformation energies of the interacting species and their interaction energies.  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