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.
The redox chemistry of [Cp*Fe(η5-As5)] ( 1 , Cp*=η5-C5Me5) has been investigated by cyclic voltammetry, revealing a redox behavior similar to that of its lighter congener [Cp*Fe(η5-P5)]. However, the subsequent chemical reduction of 1 by KH led to the formation of a mixture of novel Asn scaffolds with n up to 18 that are stabilized only by [Cp*Fe] fragments. These include the arsenic-poor triple-decker complex [K(dme)2][{Cp*Fe(μ,η2:2-As2)}2] ( 2 ) and the arsenic-rich complexes [K(dme)3]2[(Cp*Fe)2(μ,η4:4-As10)] ( 3 ), [K(dme)2]2[(Cp*Fe)2(μ,η2:2:2:2-As14)] ( 4 ), and [K(dme)3]2[(Cp*Fe)444:3:3:2:2:1:1-As18)] ( 5 ). Compound 4 and the polyarsenide complex 5 are the largest anionic Asn ligand complexes reported thus far. Complexes 2 – 5 were characterized by single-crystal X-ray diffraction, 1H NMR spectroscopy, EPR spectroscopy ( 2 ), and mass spectrometry. Furthermore, DFT calculations showed that the intermediate [Cp*Fe(η5-As5)], which is presumably formed first, undergoes fast dimerization to the dianion [(Cp*Fe)2(μ,η4:4-As10)]2−.  相似文献   

2.
The oxidation of [(Cp*Mo)2(μ,η66-P6)] ( 1 ) with halogens or halogen sources was investigated. The iodination afforded the ionic complexes [(Cp*Mo)2(μ,η33-P3)(μ,η1111-P3I3)][X] (X=I3, I) ( 2 ) and [(Cp*Mo)2(μ,η44-P4)(μ-PI2)][I3] ( 3 ), while the reaction with PBr5 led to the complexes [(Cp*Mo)2(μ,η33-P3)(μ-Br)2][Cp*MoBr4] ( 4 ) [(Cp*MoBr)2(μ,η33-P3)(μ,η1-P2Br3)] ( 5 ) and [(Cp*Mo)2(μ-PBr2)(μ-PHBr)(μ-Br)2] ( 6 ). The reaction of 1 with the far stronger oxidizing agent PCl5 was followed via time- and temperature-dependent 31P{1H} NMR spectroscopy. One of the first intermediates detected at 193 K was [(Cp*Mo)2(μ,η33-P3)(μ-PCl2)2][PCl6] ( 8 ) which rearranges upon warming to [(Cp*Mo)2(μ-PCl2)2(μ-Cl)2] ( 9 ), [(Cp*MoCl)2(μ,η33-P3)(μ-PCl2)] ( 10 ) and [(Cp*Mo)2(μ,η44-P4)(μ-PCl2)][Cp*MoCl4] ( 11 ), which could be isolated at room temperature. All complexes were characterized by single-crystal X-ray diffraction, NMR spectroscopy and their electronic structures were elucidated by DFT calculations.  相似文献   

3.
In a high‐yield one‐pot synthesis, the reactions of [Cp*M(η5‐P5)] (M=Fe ( 1 ), Ru ( 2 )) with I2 resulted in the selective formation of [Cp*MP6I6]+ salts ( 3 , 4 ). The products comprise unprecedented all‐cis tripodal triphosphino‐cyclotriphosphine ligands. The iodination of [Cp*Fe(η5‐As5)] ( 6 ) gave, in addition to [Fe(CH3CN)6]2+ salts of the rare [As6I8]2? (in 7 ) and [As4I14]2? (in 8 ) anions, the first di‐cationic Fe‐As triple decker complex [(Cp*Fe)2(μ,η5:5‐As5)][As6I8] ( 9 ). In contrast, the iodination of [Cp*Ru(η5‐As5)] ( 10 ) did not result in the full cleavage of the M?As bonds. Instead, a number of dinuclear complexes were obtained: [(Cp*Ru)2(μ,η5:5‐As5)][As6I8]0.5 ( 11 ) represents the first Ru‐As5 triple decker complex, thus completing the series of monocationic complexes [(CpRM)2(μ,η5:5‐E5)]+ (M=Fe, Ru; E=P, As). [(Cp*Ru)2As8I6] ( 12 ) crystallizes as a racemic mixture of both enantiomers, while [(Cp*Ru)2As4I4] ( 13 ) crystallizes as a symmetric and an asymmetric isomer and features a unique tetramer of {AsI} arsinidene units as a middle deck.  相似文献   

4.
The oxidation of [(Cp’’’Co)2(μ,η2 : η2-E2)2] (E=As ( 1 ), P ( 2 ); Cp’’’=1,2,4-tri(tert-butyl)cyclopentadienyl) with halogens or halogen sources (I2, PBr5, PCl5) was investigated. For the arsenic derivative, the ionic compounds [(Cp’’’Co)2(μ,η4 : η4−As4X)][Y] (X=I, Y=[As6I8]0.5 ( 3 a ), Y=[Co2Cl6-nIn]0.5 (n=0, 2, 4; 3 b ); X=Br, Y=[Co2Br6]0.5 ( 4 ); X=Cl, Y=[Co2Cl6]0.5 ( 5 )) were isolated. The oxidation of the phosphorus analogue 2 with bromine and chlorine sources yielded the ionic complexes [(Cp’’’Co)2(μ-PBr2)2(μ-Br)][Co2Br6]0.5 ( 6 a ), [(Cp’’’Co)2(μ-PCl2)2(μ-Cl)][Co2Cl6]0.5 ( 6 b ) and the neutral species [(Cp’’’Co)2(μ-PCl2)(μ-PCl)(μ,η1 : η1-P2Cl3] ( 7 ), respectively. As an alternative approach, quenching of the dications [(Cp’’’Co)2(μ,η4 : η4-E4)][TEF]2 (TEF=[Al{OC(CF3)3}4], E=As ( 8 ), P ( 9 )) with KI yielded [(Cp’’’Co)2(μ,η4 : η4-As4I)][I] ( 10 ), representing the homologue of 3 , and the neutral complex [(Cp’’’Co)(Cp’’’CoI2)(μ,η4 : η1-P4)] ( 11 ), respectively. The use of [(CH3)4N]F instead of KI leads to the formation of [(Cp’’’Co)2(μ-PF2)(μ,η2 : η1 : η1-P3F2)] ( 12 ) and 2 , thereby revealing synthetic access to polyphosphorus compounds bearing P−F groups and avoiding the use of very strong fluorinating reagents, such as XeF2, that are difficult to control.  相似文献   

5.
Using the redox-active tetrathiafulvalene tetrabenzoate (TTFTB4−) as the linker, a series of stable and porous rare-earth metal–organic frameworks (RE-MOFs), [RE93-OH)133-O)(H2O)9(TTFTB)3] ( 1-RE , where RE=Y, Sm, Gd, Tb, Dy, Ho, and Er) were constructed. The RE93-OH)133-O) (H2O)9](CO2)12 clusters within 1-RE act as segregated single-molecule magnets (SMMs) displaying slow relaxation. Interestingly, upon oxidation by I2, the S=0 TTFTB4− linkers of 1-RE were converted into S= TTFTB.3− radical linkers which introduced exchange-coupling between SMMs and modulated the relaxation. Furthermore, the SMM property can be restored by reduction in N,N-dimethylformamide. These results highlight the advantage of MOFs in the construction of redox-switchable SMMs.  相似文献   

6.
Reduction of neutral metal clusters (Co4(CO)12, Ru3(CO)12, Fe3(CO)12, Ir4(CO)12, Rh6(CO)16, {CpMo(CO)3}2, {Mn(CO)5}2) by decamethylchromocene (Cp*2Cr) or sodium fluorenone ketyl in the presence of cryptand[2.2.2] and DB‐18‐crown‐6 was studied. Nine new salts with paramagnetic Cp*2Cr+, cryptand[2.2.2](Na+), and DB‐18‐crown‐6(Na+) cations and [Co6(CO)15]2– ( 1 , 2 ), [Ru6(CO)18]2– ( 3 – 4 ) dianions, [Rh11(CO)23]3– ( 6 ) trianions, and new [Ir8(CO)18]2– ( 5 ) dianions were obtained and structurally characterized. The increase of nuclearity of clusters under reduction was shown. Fe3(CO)12 preserves the Fe3 core under reduction forming the [Fe3(CO)11]2– dianions in 7 . The [CpMo(CO)3]2 and [Mn(CO)5]2 dimers dissociate under reduction forming mononuclear [CpMo(CO)3] ( 8 ) and [Mn(CO)5] ( 9 ) anions. In all anions the increase of negative charge on metal atoms shifts the bands attributed to carbonyl C–O stretching vibrations to smaller wavenumbers in agreement with the elongation of the C–O bonds in 1 – 9 . In contrast, the M–C(CO) bonds are noticeably shortened at the reduction. Magnetic susceptibility of the salts with Cp*2Cr+ is defined by high spin Cp*2Cr+ (S = 3/2) species, whereas all obtained anionic metal clusters and mononuclear anions are diamagnetic. Rather weak magnetic coupling between S = 3/2 spins is observed with Weiss temperature from –1 to –11 K. That is explained by rather long distances between Cp*2Cr+ and the absence of effective π–π interaction between them except compound 7 showing the largest Weiss temperature of –11 K. The {DB‐18‐crown‐6(Na+)}2[Co6(CO)15]2– units in 2 are organized in infinite 1D chains through the coordination of carbonyl groups of the Co6 clusters to the Na+ ions and π–π stacking between benzo groups of the DB‐18‐crown‐6(Na+) cations.  相似文献   

7.
Heterometallic Complexes with E6 Ligands (E = P, As) The reaction of [Cp*Co(μ-CO)]2 1 with the sandwich complexes [Cp*Fe(η5-E5)] 2 a: E = P, 2 b: E = As in decalin at 190°C affords besides [CpCo2E4] 4: E = P, 7: E = As and [CpFe2P4] 5 the trinuclear complexes [(Cp*Fe)2(Cp*Co)(μ-η2-P2)(μ31:2:1-P2)2] 3 as well as [(Cp*Fe)2(Cp*Co)(μ32:2:2-As3)2] 6 . With [Mo(CO)5(thf)] 3 and 6 form in a build-up reaction the tetranuclear clusters [(Cp*Fe)2(Cp*Co)E6{Mo(CO)3}] 10: E = P, 11: E = As. 3, 6 and 11 have been further characterized by an X-ray crystal structure determination.  相似文献   

8.
By the reaction of [Cp*Fe(η5-As5)] ( I ) (Cp*=C5Me5) with main group nucleophiles, unique functionalized products with η4-coordinated polyarsenide (Asn) units (n=5, 6, 20) are obtained. With carbon-based nucleophiles such as MeLi or KBn (Bn=CH2Ph), the anionic organo-substituted polyarsenide complexes, [Li(2.2.2-cryptand)][Cp*Fe(η4-As5Me)] ( 1 a ) and [K(2.2.2-cryptand)][Cp*Fe{η4-As5(CH2Ph)}] ( 1 b ), are accessible. The use of KAsPh2 leads to a selective and controlled extension of the As5 unit and the formation of the monoanionic compound [K(2.2.2-cryptand][Cp*Fe(η4-As6Ph2)] ( 2 ). When I is reacted with [M]As(SiMe3)2 (M=Li ⋅ THF; K), the formation of the largest known anionic polyarsenide unit in [M′(2.2.2-cryptand)]2[(Cp*Fe)45443311-As20}] ( 3 ) occurred (M′=Li ( 3 a ), K ( 3 b )).  相似文献   

9.
Metalloradical species [Co2Fv(CO)4].+ ( 1 .+, Fv=fulvalenediyl) and [Co2Cp2(CO)4].+ ( 2 .+, Cp=η5‐C5H5), formed by one‐electron oxidations of piano‐stool cobalt carbonyl complexes, can be stabilized with weakly coordinating polyfluoroaluminate anions in the solid state. They feature a supported and an unsupported (i.e. unbridged) cobalt–cobalt three‐electron σ bond, respectively, each with a formal bond order of 0.5 (hemi‐bond). When Cp is replaced by bulkier Cp* (Cp*=η5‐C5Me5), an interchange between an unsupported radical [Co2Cp*2(CO)4].+ (anti‐ 3 .+) and a supported radical [Co2Cp*2(μ‐CO)2(CO)2].+ (trans‐ 3 .+) is observed in solution, which cocrystallize and exist in the crystal phase. 2 .+ and anti‐ 3 .+ are the first stable thus isolable examples that feature an unsupported metal–metal hemi‐bond, and the coexistence of anti‐ 3 .+ and trans‐ 3 .+ in one crystal is unprecedented in the field of dinuclear metalloradical chemistry. The work suggests that more stable metalloradicals of metal–metal hemi‐bonds may be accessible by using metal carbonyls together with large and weakly coordinating polyfluoroaluminate anions.  相似文献   

10.
Building upon our earlier results on the chemistry of nido-1,2-[(Cp*RuH)2B3H7] (Cp*=ɳ5-C5Me5) (nido- 1 ) with different transition metal carbonyls, we continued to investigate the reactivity with group 7 metal carbonyls under photolytic condition. Photolysis of nido- 1 with [Mn2(CO)10] led to the isolation of a trimetallic [(Cp*Ru)2{Mn(CO)3}(μ-H)(μ-CO)3(μ3-BH)] ( 2 ) cluster with a triply bridging borylene moiety. Cluster 2 is a rare example of a tetrahedral cluster having hydrido(hydroborylene) moiety. In an attempt to synthesize the Re analogue of 2 , a similar reaction was carried out with [Re2(CO)10] that yielded the trimetallic [(Cp*Ru)2{Re(CO)3}(μ-H)(μ-CO)3(μ3-BH)] ( 3 ) cluster having a triply bridging borylene unit. Along with 3 , a trimetallic square pyramid cluster [(Cp*Ru)2{Re(CO)3}(μ-H)2(μ-CO)(μ3,ɳ2-B2H5)] ( 4 ), and heterotrimetallic hydride clusters [{Cp*Ru(CO)2}-{Re(CO)4}2(μ-H)] ( 5 ) and [{Cp*Ru(CO)}{Re(CO)4}2(μ-H)3] ( 6 ) were isolated. Cluster 4 is a unique example of a M2M′B2 cluster having diboron capped Ru2Re-triangle. The hydride clusters 5 and 6 have triangular RuRe2 frameworks with one and three μ-Hs respectively. All the clusters have been characterized by using mass spectrometry, 1H, 11B{1H}, 13C{1H} NMR and IR spectroscopies analyses and the structures of clusters 2 – 6 have been unambiguously established by XRD analyses. Furthermore, to understand the electronic, structural, and bonding features of the synthesized metal-rich clusters, DFT calculations have been performed.  相似文献   

11.
The reaction of [{(η5-Me5C5)Co}2(μ-η66-toluene)] with water under different conditions leads to formation of the clusters [{(η5-Me5C5)Co}33-O)2] (1), [{(η5-Me5C5)Co}43-OH)4] (2) and [{(η5-Me5C5)Co}33-OH)4]2Co (3), whereas its reaction with hydrogen sulfide leads to [{(η5-Me5C5)Co}33-S)4]Co(μ3-S)2[(η5-Me5C5)Co]2 · Et2O (4). 1, 2 and 4 were characterized by single crystal X-ray diffraction. 1 is composed of a central Co3O2 unit, with two O2- units in an apical postion. The three cobalt atoms form a regular triangle with Co–Co distances of 2.438(2) Å, and the two oxygen atoms are located in apical positions of the triangular arrangement. The pentamethylcyclopentadienyl (Cp*) ligands are terminally bonded to the Co atoms in a η5-fashion. The Co and O atoms of 2 form a cubane-type Co4O4 cluster, with η5-bonded Cp* ligands. The central unit of 3 consists of a Co7O8 double cubane framework. Two Co4O4 cluster sharing a common corner (Co atom). Each of the other six Co atoms of the double cubane bound terminally a Cp* ligand. 4 is composed of a [{(η5-Me5C5)Co }23-S)4](μ3-S)2[(η5-Me5C5)Co]2 and an ether molecule. 4 contains a central Co4S4 cubane-like unit. One of the four Co atoms is bonded via two μ3-S atoms to two additional (η5-Cp*Co) units. The other three Co atoms are η5-coordinated to a Cp* ligand.  相似文献   

12.
Zincocene Cp*2Zn reacts with carbodiimides C(NR)2 with insertion into the Zn–Cp* bond and formation of [(Cp*C(NR)2]2Zn [R = Et ( 1 ), iPr ( 2 ), Cy ( 3 )]. In addition, the reaction of Cp*2Zn with CS2 under dry conditions gives (Cp*CS2)2Zn ( 4 ), whereas in the presence of a small amount of water [Zn44‐O)(S2CCp*)6] ( 5 ) is obtained. Compounds 1 – 4 were characterized by NMR (1H, 13C) and IR spectroscopy as well as elemental analysis and single‐crystal X‐ray diffraction ( 2 – 4 , 5 of poor quality). The solid‐state structure of 5 is comparable to the carboxylate complex previously obtained from the reaction of Cp*2Zn with CO2.  相似文献   

13.
A route to directly access mixed Al–Fe polyphosphide complexes was developed. The reactivity of pentaphosphaferrocene, [Cp*Fe(η5‐P5)] (Cp*=C5Me5), with two different low‐valent aluminum compounds was investigated. The steric and electronic environment around the [AlI] centre are found to be crucial for the formation of the resulting Al–Fe polyphosphides. Reaction with the sterically demanding [Dipp‐BDIAlI] (Dipp‐BDI={[2,6‐iPr2C6H3NCMe]2CH}?) resulted in the first Al‐based neutral triple‐decker type polyphosphide complex. For [(Cp*AlI)4], an unprecedented regioselective insertion of three [Cp*AlIII]2+ moieties into two adjacent P?P bonds of the cyclo‐P5 ring of [Cp*Fe(η5‐P5)] was observed. The regioselectivity of the insertion reaction could be rationalized by isolating an analogue of the reaction intermediate stabilized by a strong σ‐donor carbene.  相似文献   

14.
Triply‐bridging bis‐{hydrido(borylene)} and bis‐borylene species of groups 6, 8 and 9 transition metals are reported. Mild thermolysis of [Fe2(CO)9] with an in situ produced intermediate, generated from the low‐temperature reaction of [Cp*WCl4] (Cp*=η5‐C5Me5) and [LiBH4?THF] afforded triply‐bridging bis‐{hydrido(borylene)}, [(μ3‐BH)2H2{Cp*W(CO)2}2{Fe(CO)2}] ( 1 ) and bis‐borylene, [(μ3‐BH)2{Cp*W(CO)2}2{Fe(CO)3}] ( 2 ). The chemical bonding analyses of 1 show that the B?H interactions in bis‐{hydrido (borylene)} species is stronger as compared to the M?H ones. Frontier molecular orbital analysis shows a significantly larger energy gap between the HOMO‐LUMO for 2 as compared to 1 . In an attempt to synthesize the ruthenium analogue of 1 , a similar reaction has been performed with [Ru3(CO)12]. Although we failed to get the bis‐{hydrido(borylene)} species, the reaction afforded triply‐bridging bis‐borylene species [(μ3‐BH)2{WCp*(CO)2}2{Ru(CO)3}] ( 2′ ), an analogue of 2 . In search for the isolation of bridging bis‐borylene species of Rh, we have treated [Co2(CO)8] with nido‐[(RhCp*)2(B3H7)], which afforded triply‐bridging bis‐borylene species [(μ3‐BH)2(RhCp*)2Co2(CO)4(μ‐CO)] ( 3 ). All the compounds have been characterized by means of single‐crystal X‐ray diffraction study; 1H, 11B, 13C NMR spectroscopy; IR spectroscopy and mass spectrometry.  相似文献   

15.
A series of new heteromultinuclear FeI/RuII clusters are described. The complexes (η6-arene)RuFe2S2(CO)6 (arene = p-cymene 1 , C6Me6 2 ) and Fe2[μ-S (Cp*Ru)(CO)2]2(CO)6 (Cp* = η5-C5Me5) ( 3 ) were prepared by the reduction reactions of (μ-S)2Fe2(CO)6 with 2 equiv of LiHBEt3, followed by treatment (μ-SLi)2Fe2(CO)6 with ruthenium-arene complexes Ru2(μ-Cl)2Cl2(η6-arene)2 or Cp*Ru (CO)2Cl in 22–33% yields. Further reactions of 1 and 2 with 1 equiv of triphenylphosphine in the presence of the decarbonylating agent Me3NO·2H2O, afforded the corresponding monophosphine-substituted FeI/RuII complexes (η6-arene)RuFe2S2(CO)5(Ph3P) (arene = p-cymene 4 , C6Me6 5 ) in 75% and 78% yields. While treatment of parent complex 1 or 2 with 1 equiv of diphosphine Ph2PCH2PPh2 (dppm) in xylene at reflux temperature resulted in the formation of the diphosphine-bridged RuFe2S2(CO)9 derivate RuFe2S2(CO)7(dppm) ( 6 ). The possible pathway for the formation was proposed. Two isomers of novel macrocyclic complexes involve the (η6-arene) Ru-bridged quadruple-butterfly Fe/S clusters [{μ-S (CH2)3S-μ}{(μ-CS2)Fe2(CO)6}2]2[(η6-arene)Ru]2 (arene = p-cymene 7a and 7b , C6Me6 8a and 8b ) were isolated by reactions of two μ-CS2-containing dianion [{μ-S (CH2)3S-μ}{(μ-S=CS)Fe2(CO)6}2]2− with [Ru2(μ-Cl)2Cl2(η6-arene)2], in which the propylene groups are attached to two S atoms by ee and ea types of bonds respectively. All the new complexes 1 – 8 have been characterized by elemental analysis, spectroscopy, and particularly for 1 – 6 , 7b and 8a by X-ray crystallography. In addition, the electrochemical properties of representative complexes 1 – 4 and 6 have been investigated.  相似文献   

16.
Sodium nitrosylcarbonyliron reacts with methylcyclopentadienylcarbonylmetal(Mo orW)chloride in CH_3OH/THF at room temperature to give CpMo(CO)_2NO(1a)(Cp=η~5-CH_3C_5H_4)or CpW(CO)_2NO(1b),[CpMo(CO)_3]_2(2a)or[CpW(CO)_3]_2(2b),and CpMo(μ3-NH)(μ2-NO)-(μ2-CO)Fe_2(CO)_6(3a)or CpW(μ3-NH)(μ2-NO)(μ2-CO)Fe_2(CO)_6(3b),respectively.Complexes1a,1b,3a and 3b were analyzed by IR,NMR,MS and elemental analyses,and the crystalstructures of 1b,3a and 3b were determined by X-ray diffraction method.The new clusters 3aand 3b have μ3-NH ligands which were formed by redaction of NO in the synthetic reactions.  相似文献   

17.
Synthesis, bonding and chemistry of mono- and bimetallic complexes supported by chelating thiolato ligands have been established. Treatment of [Cp*VCl2]3 ( 1 ) with [LiBH4 ⋅ THF] followed by the addition of ethane-1,2-dithiol led to the formation of an EPR active bimetallic vanadium thiolato complex [(Cp*V){μ-(SCH2CH2S)-κ2S,S′)2{V(SCH2CH2S-SH)}] ( 2 ). In complex 2 , two ethane-1,2-dithiolato ligands are symmetrically coordinated to two vanadium atoms through μ-S atoms. Interestingly, when similar reactions were carried out with heavier group 5 metal precursors, such as [Cp*NbCl4] ( 3 a ), it afforded monometallic thiolato complex [Cp*Nb(SCH2CH2S)(SCH2CH2S−CH2S)] ( 4 a ). On the other hand, the Ta-analogue [Cp*TaCl4] ( 3 b ) yielded thiolato species [Cp*Ta(SCH2CH2S)(SCH2CH2S−CH2S)] ( 4 b ) and [Cp*Ta(SCH2CH2S) (SCH2CH2S−S)] ( 5 ). In complexes 4 a and 4 b , one ethane-1,2-dithiolato and one trithiolato ligand are coordinated to Nb and Ta centers, respectively. Whereas, in complex 5 , one ethane-1,2-dithiolato and one 2-disulfanylethanethiolato is coordinated to the Ta center. Moreover, the photolytic reaction of 5 with [Mo(CO)5 ⋅ THF] yielded heterobimetallic thiolato complex [(Cp*Ta){μ-(SCH2CH2S)-κ2S,S′}{μ-(SCH2CH2S−CH2(CH3)S)κ2S′′ : κ1S-′′′′ : κ1S′′′′′}{Mo(CO)3}] ( 6 ). All the complexes have been characterized by multinuclear NMR spectroscopy and single crystal X-ray diffraction studies. Further, computational analyses were performed to provide an insight into the bonding of these complexes.  相似文献   

18.
The hexachalcogenodistannates K6[SnIII2Se6] or Li4[SnIV2Te6]·8en were recently reported to simultaneously act as mild oxidants and chalcogenide sources in reactions with CoCl2/LiCp* (Cp* = pentamethylcyclopentadienide) while the Sn—E (E = Se, Te) fragment is not kept in the products, e.g. [(Cp*Co)3(μ3‐Se)2], [(Cp*Co)3(μ3‐Se)2][Cl2Co(μ2‐Cl)2Li(thf)2] or [(Cp*Co)4(μ3‐Te)4]. In search of related reagents with possibly different reaction behavior, we isolated and crystallographically characterized isotypic compounds [enH]4[SnIV2Se6]�en ( 1 ), and [enH]4[SnIV2Te6en ( 2 ) (en = 1, 2‐diaminoethane), that result from an uncommon disproportion/re‐arrangement reaction: distannate(III) K6[Sn2E6] (E = Se, Te) was reacted with en·2HCl to yield 1 or 2 under disproportion of SnIII to SnII and SnIV. Another pathway was necessary to synthesize the respective but solvent‐free thiostannate [enH]4 [SnIV2S6] ( 3 ), since the phase “K6[Sn2S6]” is unknown. This second method started out from SnCl4·2THF and S(SiMe3)2 in en solution. However, using E(SiMe3)2 (E = Se, Te) instead of S(SiMe3)2, 1 and 2 are also obtained this way. 1—3 are the first chalcogenostannates that exhibit exclusively [enH]+ counterions. The compounds were characterized by means of X‐ray crystallography and NMR spectroscopy. They seem to be suitable for reactions towards group 8‐10 metal complexes. Preliminary experiments indicate that the binary anions 1 — 3 coordinated by 1‐aminoethylammonium ions react more slowly compared to the anionic phases tested until now.  相似文献   

19.
Reaction of [1,2‐(Cp*RuH)2B3H7] ( 1 ; Cp*=η5‐C5Me5) with [Mo(CO)3(CH3CN)3] yielded arachno‐[(Cp*RuCO)2B2H6] ( 2 ), which exhibits a butterfly structure, reminiscent of 7 sep B4H10. Compound 2 was found to be a very good precursor for the generation of bridged borylene species. Mild pyrolysis of 2 with [Fe2(CO)9] yielded a triply bridged heterotrinuclear borylene complex [(μ3‐BH)(Cp*RuCO)2(μ‐CO){Fe(CO)3}] ( 3 ) and bis‐borylene complexes [{(μ3‐BH)(Cp*Ru)(μ‐CO)}2Fe2(CO)5] ( 4 ) and [{(μ3‐BH)(Cp*Ru)Fe(CO)3}2(μ‐CO)] ( 5 ). In a similar fashion, pyrolysis of 2 with [Mn2(CO)10] permits the isolation of μ3‐borylene complex [(μ3‐BH)(Cp*RuCO)2(μ‐H)(μ‐CO){Mn(CO)3}] ( 6 ). Both compounds 3 and 6 have a trigonal‐pyramidal geometry with the μ3‐BH ligand occupying the apical vertex, whereas 4 and 5 can be viewed as bicapped tetrahedra, with two μ3‐borylene ligands occupying the capping position. The synthesis of tantalum borylene complex [(μ3‐BH)(Cp*TaCO)2(μ‐CO){Fe(CO)3}] ( 7 ) was achieved by the reaction of [(Cp*Ta)2B4H8(μ‐BH4)] at ambient temperature with [Fe2(CO)9]. Compounds 2 – 7 have been isolated in modest yield as yellow to red crystalline solids. All the new compounds have been characterized in solution by mass spectrometry; IR spectroscopy; and 1H, 11B, and 13C NMR spectroscopy and the structural types were unequivocally established by crystallographic analysis of 2 – 6 .  相似文献   

20.
The speciation of compounds [Cp*2M2O5] (M=Mo, W; Cp*=pentamethylcyclopentadienyl) in different protic and aprotic polar solvents (methanol, dimethyl sulfoxide, acetone, acetonitrile), in the presence of variable amounts of water or acid/base, has been investigated by 1H NMR spectrometry and electrical conductivity. Specific hypotheses suggested by the experimental results have been further probed by DFT calculations. The solvent (S)‐assisted ionic dissociation to generate [Cp*MO2(S)]+ and [Cp*MO3]? takes place extensively for both metals only in water/methanol mixtures. Equilibrium amounts of the neutral hydroxido species [Cp*MO2(OH)] are generated in the presence of water, with the relative amount increasing in the order MeCN≈acetone<MeOH<DMSO. Addition of a base (Et3N) converts [Cp*2M2O5] into [Et3NH]+[Cp*MO3]?, for which the presence of a N? H???O?M interaction is revealed by 1H NMR spectroscopy in comparison with the sodium salts, Na+[Cp*MO3]?. These are fully dissociated in DMSO and MeOH, but display a slow equilibrium between free ions and the ion pair in MeCN and acetone. Only one resonance is observed for mixtures of [Cp*MO3]? and [Cp*MO2(OH)] because of a rapid self‐exchange. In the presence of extensive ionic dissociation, only one resonance is observed for mixtures of the cationic [Cp*MO2(S)]+ product and the residual undissociated [Cp*2M2O5] because of a rapid associative exchange via the trinuclear [Cp*3M3O7]+ intermediate. In neat methanol, complex [Cp*2W2O5] reacts to yield extensive amounts of a new species, formulated as the mononuclear methoxido complex [Cp*WO2(OMe)] on the basis of the DFT study. An equivalent product is not observed for the Mo system. The addition of increasing amounts of water results in the rapid decrease of this product in favor of [Cp*2W2O5] and [Cp*WO2(OH)].  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