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 46 毫秒
1.
Synthesis of Carboxylate Substituted Rhenium Gold Metallatetrahedranes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H, Me, CF3, Ph, 3,4-(OMe)2C6H3) The reaction of the in situ prepared salt Li[Re2(μ-H)(μ-PCy2)(CO)7(ax-C(Ph)O)] ( 2 ) with 1,5 equivalents of monocarboxylic acid RCOOH (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e ) in tetrahydrofruan (THF) solution at 60 °C gives within 4 h under release of benzaldehyde (PhCHO) the η1-carboxylate substituted dirhenium salt Li[Re2(μ-H)(μ-PCy2)(CO)71-OC(R)O)] (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e )) in almost quantitative yield. The lower the pKa value of the respective carboxylic acid the faster the reaction proceeds. It was only in the case of CF3COOH possible to prove the formation of the hydroxycarbene complex Re2(μ-H)(μ-PCy2)(CO)7(=C(Ph)OH) ( 5 ) prior to elimination of PhCHO. The new compounds 4 a–4 e were only characterized by 31P NMR and ν(CO) IR spectroscopy as they are only stable in solution. They are converted with two equivalents of BF4AuPPh3 at 0 °C in a so-called cluster expansion reaction into the heterometallic metallatetrahedrane complexes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H ( 7 a ), Me ( 7 b ), CF3 ( 7 c ), Ph ( 7 d ), 3,4-(OMe)2C6H3 ( 7 e )) (yield 47–71% ). The expected precursor complexes of 7 a–7 e Li[Re2(AuPPh3)(μ-PCy2)(CO)71-OC(R)O] ( 8 ) were not detected by NMR and IR spectroscopy in the course of the reaction. Their existence was retrosynthetically proved by the reaction of 7 b with an excess of the chelating base TBD (1,5,7-Triazabicyclo[4.4.0]dec-5-en) forming [(TBD)xAuPPh3][Re2(AuPPh3)(μ-PCy2)(CO)71-OC(Me)O] ( 8 b ) in solution. The η1-bound carboxylate ligand in 7 a–7 e can photochemically be converted into a μ-bound ligand in Re2(AuPPh3)2(μ-PCy2)(μ-OC(R)O)(CO)6 (R = H ( 9 a ), Me ( 9 b ), CF3 ( 9 c ), Ph ( 9 d ), 3.4-(MeO)2C6H3 ( 9 e )) under release of one equivalent CO. All isolated cluster complexes were characterized and identified by the following analytical methods: elementary analysis, NMR (1H, 31P) spectroscopy, ν(CO) IR spectroscopy and in the case of 7 d and 9 b by X-ray structure analysis.  相似文献   

2.
Treatment of the electronically unsaturated cluster [(μ-H)Os3(CO)8{Ph2PCH2P(Ph)C6H4}] (1) with primary phosphines PPhH2 and PCyH2 gives the phosphido bridged compounds [(μ-H)Os3(CO)8(μ-PPhH)(μ-dppm)] (2) and [(μ-H)Os3(CO)8(μ-PCyH)(μ-dppm)] (3), respectively, by P-H bond activation of the phosphines and demetallation of the phenyl ring of the diphosphine ligand. Thermolysis of 2 and 3 in refluxing octane at 128 °C results in the formation of the phosphinidene compounds [(μ-H)2Os3(CO)73-PPh)(μ-dppm)] (4) and [(μ-H)2Os3(CO)73-PCy)(μ-dppm)] (5), respectively, by further P-H bond cleavage of the phosphido groups. All the compounds have been characterized by infrared, 1H NMR, 31P{1H} NMR and mass spectroscopic data together with single-crystal X-ray diffraction studies for 4. Compound 4 consists of a triangular cluster of osmium atoms with a symmetrically capped phosphinidene ligand and a bridging dppm ligand.  相似文献   

3.
The preparations of the binuclear hydrido-bridged cations [(terdentate ligand)Pd(μ-H)Pd(terdentate ligand)]+ from [(terdentate ligand)Pd(acetone)]+ and NaO2CH and [(terdentate ligand)Pd(μ-H)Pt(terdentate ligand)]+ from [(terdentate ligand)Pd(acetone)]+ and [(terdentate ligand)PtH] (terdentate ligand = 2,6-(Ph2PCH2)2C6H3) are reported. The preparation of the cation [(terdentate ligand)Pt(μ-H)Pt(terdentate ligand)]+ is also reported.  相似文献   

4.
The new complexes [(η3-Me2CCMeCH2)Pd{η2-Ph2P(S)CHP(S)Ph2] (1), [(η3-Me2CCMeCH2)Pd{η2-OC(CF3) CHCO(C4H3S)}] (2) and [(η3-CH2CMeCH2)Pd{η2-OC(CF3)CHCO(C4H3S)}] (3) have been synthesized by reacting [(η3-allyl)Pd(μ-Cl)]2 with Ph2P(S)CH2P(S)Ph2 and OC(CF3)CH2CO(C4H3S) in the presence of base. All have been characterized by elemental analysis, FT-IR, 1H-n.m.r and FAB-mass spectroscopy. Spectroscopic studies suggest that both ligands are bidentate, forming six-membered Pd-S-P-C-P-S and Pd-O-C-C-C-O palladacycles, the η3-allyl group completing the coordination sphere.  相似文献   

5.
Alkylzinc alkoxides, [RZnOR′]4, have received much attention as efficient precursors of ZnO nanocrystals (NCs), and their “Zn4O4” heterocubane core has been regarded as a “preorganized ZnO”. A comprehensive investigation of the synthesis and characterization of a new family of tert‐butyl(tert‐butoxy)zinc hydroxides, [(tBu)4Zn43‐OtBu)x3‐OH)4?x], as model single‐source precursors of ZnO NCs is reported. The direct reaction between well‐defined [tBuZnOH]6 ( 16 ) and [tBuZnOtBu]4 ( 24 ) in various molar ratios allows the isolation of new mixed cubane aggregates as crystalline solids in a high yield: [(tBu)4Zn43‐OtBu)33‐OH)] ( 3 ), [(tBu)4Zn43‐OtBu)23‐OH)2] ( 4 ), [(tBu)4Zn43‐OtBu)(μ3‐OH)3] ( 5 ). The resulting products were characterized in solution by 1H NMR and IR spectroscopy, and in the solid state by single‐crystal X‐ray diffraction. The thermal transformations of 2 – 5 were monitored by in situ variable‐temperature powder X‐ray diffraction and thermogravimetric measurements. The investigation showed that the Zn?OH groups appeared to be a desirable feature for the solid‐state synthesis of ZnO NCs that significantly decreased the decomposition temperature of crystalline precursors 3 – 5 .  相似文献   

6.
The phosphine oxide complexes [GaX3(Me3PO)] and [(GaX3)2{μ-o-C6H4(CH2P(O)Ph2)2}] have been prepared and characterised by microanalysis, IR and multinuclear NMR (1H, 13C{1H}, 31P{1H} and 71Ga) spectroscopy. The structures of [GaCl3(Me3PO)], [(GaBr3)2{μ-o-C6H4(CH2P(O)Ph2)2}] and of the ionic product [GaI2(Me3PO)2][GaI4] have been determined and show that the Lewis acidity of the gallium halides towards phosphinoyl ligands diminishes as the halogen becomes heavier. The [GaX3(Ph3E)] (X = Cl, Br or I; E = P or As) and [(GaX3)2{μ-o-C6H4(CH2PPh2)2}] (X = Br or I) have been prepared and their structural and spectroscopic properties compared with those of the phosphinoyl complexes. The results, and competitive solution NMR studies, show that Ga(III) binds the hard R3PO in preference to the softer phosphine or arsine ligands. Hydrolysis of gallium(III) phosphines is shown to lead to [R3PH][GaX4], but in contrast to some other p-block halides, GaX3 do not promote air-oxidation of R3P to R3PO.  相似文献   

7.
The synthesis, structural characterization, and reactivity of new bridged borylene complexes are reported. The reaction of [{Cp*CoCl}2] with LiBH4 ? THF at ?70 °C, followed by treatment with [M(CO)3(MeCN)3] (M=W, Mo, and Cr) under mild conditions, yielded heteronuclear triply bridged borylene complexes, [(μ3‐BH)(Cp*Co)2(μ‐CO)M(CO)5] ( 1 – 3 ; 1 : M=W, 2 : M=Mo, 3 : M=Cr). During the syntheses of complexes 1 – 3 , capped‐octahedral cluster [(Cp*Co)2(μ‐H)(BH)4{Co(CO)2}] ( 4 ) was also isolated in good yield. Complexes 1 – 3 are isoelectronic and isostructural to [(μ3‐BH)(Cp*RuCO)2(μ‐CO){Fe(CO)3}] ( 5 ) and [(μ3‐BH)(Cp*RuCO)2(μ‐H)(μ‐CO){Mn(CO)3}] ( 6 ), with a trigonal‐pyramidal geometry in which the μ3‐BH ligand occupies the apical vertex. To test the reactivity of these borylene complexes towards bis‐phosphine ligands, the room‐temperature photolysis of complexes 1 – 3 , 5 , 6 , and [{(μ3‐BH)(Cp*Ru)Fe(CO)3}2(μ‐CO)] ( 7 ) was carried out. Most of these complexes led to decomposition, although photolysis of complex 7 with [Ph2P(CH2)nPPh2] (n=1–3) yielded complexes 9 – 11 , [3,4‐(Ph2P(CH2)nPPh2)‐closo‐1,2,3,4‐Ru2Fe2(BH)2] ( 9 : n=1, 10 : n=2, 11 : n=3). Quantum‐chemical calculations by using DFT methods were carried out on compounds 1 – 3 and 9 – 11 and showed reasonable agreement with the experimentally obtained structural parameters, that is, large HOMO–LUMO gaps, in accordance with the high stabilities of these complexes, and NMR chemical shifts that accurately reflected the experimentally observed resonances. All of the new compounds were characterized in solution by using mass spectrometry, IR spectroscopy, and 1H, 13C, and 11B NMR spectroscopy and their structural types were unequivocally established by crystallographic analysis of complexes 1 , 2 , 4 , 9 , and 10 .  相似文献   

8.
Reaction of the parent complex (μ-PDT)Fe2-(CO)6 (A) (PDT = 1,3-SCH2CH2CH2S2?) with the bidentate N/P ligand [(Ph2P)2N(C6H4Cl-p)] in the presence of Me3NO as decarbonylating agent produced an unexpected iron–sulfur complex [(μ-PDT)Fe2(CO)5{PPh2(NHC6H4Cl-1,4)}] (1). Extending this chemistry further, two similar complexes [(μ-PDT)Fe2(CO)5{PPh2(NHC6H4NO2-1,4)}] (2) and [(μ-PDT)Fe2(CO)5{PPh2(NHC6H4CO2Et-1,4)}] (3) could be prepared from the simple substitution reactions of the precursor A with the monodentate N/P ligands Ph2P(NHC6H4NO2-1,4) and Ph2P(NHC6H4CO2Et-1,4), respectively. These new complexes, which can be considered as active site models of [FeFe] hydrogenases, have been characterized by elemental analysis, FTIR, and NMR (1H, 13C, 31P) spectroscopies, as well as by X-ray crystallography for complex 1.  相似文献   

9.
Continued exploration of the coordination behavior of derivatives of 2-benzophenone-based ligands with metal alkoxides ([M(OR)4]) was undertaken from the reaction of 2-(2-hydroxy-4-methoxybenzoyl)benzoic acid (H2-OBzA) with a series of Group 4 precursors. The products of these reactions were identified as: [(OR)2Ti(μ-(c,c-OBzA))]2 (OR?=?OCHMe2 (OPri; 1 ?2tol); OCMe3 (OBut; 2 ?THF); OCH2CMe3 (ONep; 3)), [[(OPri)3Ti(μ-OPri)Ti(OPri)2]2(μ-(μc,μ-OBzA))2]2 (4), [(ONep)3Zr(μ-ONep)2Zr(ONep)2]2(μ-(c,μ-OBzA)2) (5 ?tol), [(py)(OBut)3Zr]2(μ-(c,c-OBzA)) (6), [(OBut)2Hf(μ-OBut)]2(μ-(c,η1-OBzA)) (7) where ‘c’?=?chelating or η2; ‘μ’?=?bridging or η11(O,O’); and μc?=?bridging chelating or η11(O,O’); η2?:?η1. The metal centers for each of these compounds adopt a pseudo-octahedral geometry employing the OBzA ligand in numerous binding modes. The different functional oxygens (carboxylate, hydroxyl, and carbonyl) were employed in a variety of coordination modes for 1–7. The complexity of these OBzA-modified compounds is driven by a combination of the coordination behavior of the OBzA moieties, the size of the metal cation, and the pendant chain of the OR ligand. Solution NMR indicates a complex structure exists in solution that was considered to be consistent with the solid-state structure.  相似文献   

10.
The complex [(μ-H)Os3(μ-OCNMe2)(CO)9{P(CH2CHCH2)Ph2}] derived from the replacement of a lightly-stabilizing NMe3 ligand in [(μ-H)Os3(μ-OCNMe2)(CO)9(NMe3)] by allyldiphenylphosphine molecule was physico-chemically and X-ray structurally characterized and served further as a metal cluster monomer to be immobilized on a polymer surface. The copolymerization of this cluster monomer with styrene was studied. It was found that the cluster molecule, when copolymerized, maintains its integrity and the overall structure. The cluster copolymers obtained have been tested as catalysts for oxidative dehydrogenation and cyclohexenol hydroxylation reactions.  相似文献   

11.
An unsymmetric bidentate ligand (3-methyl-2-pyridyl)diphenylphosphane (P(Mepy)Ph2) is able to react with various tetranuclear transition metal clusters such as HRuCo3(CO)12, HRuRh3(CO)12 and Rh4(CO)12. The synthesis and crystal structures of HRuCo3(CO)10(P(Mepy)Ph2) (1), HRuRh3(CO)10(P(Mepy)Ph2) (2), RuRh2(CO)9(P(Mepy)Ph) (3) and Rh6(CO)14(P(Mepy)Ph2) (4) are described. In 1, 2 and 4 the phosphane ligand replaces the carbonyls and acts as a bridging bidentate P-N group. The formation of 3 includes degradation of both the metal cluster core and the ligand itself. One of the P-C bonds in the ligand is cleaved and the ligand caps a metal triangle with a bridging phosphido group together with the nitrogen donor. The reaction between dinuclear Rh2(CO)4Cl2 and P(Mepy)Ph2 gives a binuclear Rh2(μ-CO)Cl2(P(Mepy)Ph2)2 (5) with bridging ligands in a head-to-tail arrangement. The crystal structure is also given.  相似文献   

12.
The ruthenium cluster Ru3(CO)12 reacts with the diphosphine ligand 3,4-bis(diphenylphosphino)-5-methoxy-2(5H)-furanone (bmf) in refluxing toluene to furnish the donor–acceptor compound Ru2(CO)2(bmf) as a 1?:?1 mixture of diastereomers. Photolysis of Ru2(CO)2(bmf) using 366?nm light leads to the oxidative cleavage of a P–C bond and formation of the phosphido-bridged complex Ru2(CO)6[μ-C=C(PPh2)C(O)OCH(OMe)](μ-PPh2). The regioselective Ph2P–C(furanone ring) bond activation attendant upon optical excitation is traced to the phosphine group that was β to the furanone carbonyl group, as established by X-ray analysis of one of the diastereomers of Ru2(CO)6[μ-C=C(PPh2)C(O)OCH(OMe)](μ-PPh2). Both diruthenium products have been fully characterized in solution by IR and NMR (1H and 31P) spectroscopies and elemental analyses. The observed regioselectivity associated with the P–C bond activation in Ru2(CO)2(bmf) is discussed with respect to the chemistry of other bmf-substituted compounds prepared by our groups.  相似文献   

13.
The tetravalent platinum stiboranyl complex [(o‐(Ph2P)C6H4)2(o‐C6Cl4O2)Sb]PtCl2Ph ( 2 ) has been synthesized by reaction of [(o‐(Ph2P)C6H4)2SbClPh]PtCl ( 1 ) with o‐chloranil. In the presence of fluoride anions, the stiboranyl moiety of 2 displays non‐innocent behavior and is readily converted into a fluorostiborane unit. This transformation, which is accompanied by elimination of a chloride ligand from the Pt center, results in the formation of [(o‐(Ph2P)C6H4)2(o‐C6Cl4O2)SbF]PtClPh ( 3 ). Structural, spectroscopic, and computational studies show that the conversion of 2 into 3 is accompanied by a cleavage of the covalent Pt? Sb bond present in 2 and formation of a longer and weaker Pt→Sb interaction in 3 . These results show that this new Pt–Sb platform supports the fluoride‐induced metamorphosis of a stiboranyl X ligand into a stiborane Z ligand.  相似文献   

14.
UV irradiation of the diphenylchalcogenides Ph2Se2 or Ph2Te2 in the presence of [(η5-MeCp)Mo(CO)3]2 induces rapid reaction to give the double μ-EPh bridged compounds [(η5-MeCp)Mo(CO)2(μ-EPh)]2. Subsequent decarbonylation by mild thermolysis in vacuo gives [(η5-MeCp)Mo(CO)(μ-SePh)]2 or [(η5-MeCp)Mo(CO)(μ-TePh)2 in good yields. The new compounds were characterized by elemental analysis, infrared and mass spectra. The mixed Se/ Te bridged complex [(η5-MeCp)Mo2(Co)4(μ-SePh)(μ-TePh)] was not obtained by UV irradiation of [(η5-MeCp)Mo(CO)3]2 in the presence of a mixture of Ph2Se2 and Ph2Te2.  相似文献   

15.
The reaction of [CpRuCl(PPh3)2] (Cp=cyclopentadienyl) and [CpRuCl(dppe)] (dppe=Ph2PCH2CH2PPh2) with bis‐ and tris‐phosphine ligands 1,4‐(Ph2PC≡C)2C6H4 ( 1 ) and 1,3,5‐(Ph2PC≡C)3C6H3 ( 2 ), prepared by Ni‐catalysed cross‐coupling reactions between terminal alkynes and diphenylchlorophosphine, has been investigated. Using metal‐directed self‐assembly methodologies, two linear bimetallic complexes, [{CpRuCl(PPh3)}2(μ‐dppab)] ( 3 ) and [{CpRu(dppe)}2(μ‐dppab)](PF6)2 ( 4 ), and the mononuclear complex [CpRuCl(PPh3)(η1‐dppab)] ( 6 ), which contains a “dangling arm” ligand, were prepared (dppab=1,4‐bis[(diphenylphosphino)ethynyl]benzene). Moreover, by using the triphosphine 1,3,5‐tris[(diphenylphosphino)ethynyl]benzene (tppab), the trimetallic [{CpRuCl(PPh3)}33‐tppab)] ( 5 ) species was synthesised, which is the first example of a chiral‐at‐ruthenium complex containing three different stereogenic centres. Besides these open‐chain complexes, the neutral cyclic species [{CpRuCl(μ‐dppab)}2] ( 7 ) was also obtained under different experimental conditions. The coordination chemistry of such systems towards supramolecular assemblies was tested by reaction of the bimetallic precursor 3 with additional equivalents of ligand 2 . Two rigid macrocycles based on cis coordination of dppab to [CpRu(PPh3)] were obtained, that is, the dinuclear complex [{CpRu(PPh3)(μ‐dppab)}2](PF6)2 ( 8 ) and the tetranuclear square [{CpRu(PPh3)(μ‐dppab)}4](PF6)4 ( 9 ). The solid‐state structures of 7 and 8 have been determined by X‐ray diffraction analysis and show a different arrangement of the two parallel dppab ligands. All compounds were characterised by various methods including ESIMS, electrochemistry and by X‐band ESR spectroscopy in the case of the electrogenerated paramagnetic species.  相似文献   

16.
Abstract

Treatment of Ph2Te with aqueous hydrogen chloride under air in refluxing THF gave trans-[(Ph2Te)(μ-Cl)2]n (1), whereas interaction of Ph2Te with ammonium chloride under similar condition afforded cis-[(Ph2Te)(μ-Cl)2]n (2). Reaction of (p-MeC6H4)2Te and bromine in refluxing THF resulted in formation of a discrete complex [(p-MeC6H4)2TeBr2] (3) with a step-like tetrameric structure, which further reacted with sodium hydroxide in refluxing THF to give a dinuclear tellurium oxide [{(p-Me-C6H4)2TeBr}2(μ-O)] (4) with a bridging oxygen atom. Complexes 1–4 have been spectroscopically characterized and their crystal structures have been established by X-ray crystallography.  相似文献   

17.
Activation of Carbon Disulfide on Triruthenium Clusters: Synthesis and X‐Ray Crystal Structure Analysis of [Ru3(CO)4(μ‐PCy2)2(μ‐Ph2PCH2PPh2)(μ3‐S){μ3‐η2‐CSC(S)S}] [Ru3(CO)4(μ‐H)3(μ‐PCy2)3(μ‐dppm)] ( 2 ) (dppm = Ph2PCH2PPh2) reacts with CS2 at room temperature and yields the open 50 valence electron cluster [Ru3(CO)4(μ‐PCy2)2(μ‐dppm)(μ3‐S){μ3‐η2‐CSC(S)S}] ( 3 ) containing the unusual μ3‐η2‐C2S3 mercaptocarbyne ligand. Compound 3 was characterized by single crystal X‐ray structure analysis.  相似文献   

18.
Reaction of Co2(CO)8 and 1,3‐propanedithiol in a 1:1 molar ratio in toluene affords a novel tetracobalt complex, [(μ2‐pdt)23‐S)Co4(CO)6] (pdt=‐SCH2CH2CH2S‐, 1 ), which possesses some of the structural features of the active site of [FeFe]‐hydrogenase. Carbonyl displacement reaction of complex 1 in the presence of mono‐ or diphosphine ligands leads to the formation of [(μ2‐pdt)23‐S)Co4(CO)5(PCy3)] ( 2 ) and [(μ2‐pdt)23‐S)Co4(CO)4(L)] [L=Ph2PCH?CHPPh2, 3 ; Ph2PCH2N(Ph)CH2PPh2, 4 ; Ph2PCH2N(iPr)CH2PPh2, 5 ]. Complexes 1 – 5 have been fully characterized by spectroscopy and single‐crystal X‐ray diffraction studies. Cyclic voltammetry has revealed that complexes 1 – 5 show a reversible first reduction wave and are active for electrocatalytic proton reduction in the presence of CF3COOH. Protonation reactions have been monitored by 31P and 1H NMR and infrared spectroscopies, which revealed the formation of different protonated species. The mono‐reduced species of 1 – 5 have been spectroscopically characterized by EPR and spectro‐electro‐infrared techniques.  相似文献   

19.
[ReNCl2(PPh3)2] and [ReNCl2(PMe2Ph)3] react with the N‐heterocyclic carbene (NHC) 1,3,4‐triphenyl‐1,2,4‐triazol‐5‐ylidene (HLPh) under formation of the stable rhenium(V) nitrido complex [ReNCl(HLPh)(LPh)], which contains one of the two NHC ligands with an additional orthometallation. The rhenium atom in the product is five‐coordinate with a distorted square‐pyramidal coordination sphere. The position trans to the nitrido ligand is blocked by one phenyl ring of the monodentate HLPh ligand. The Re–C(carbene) bond lengths of 2.072(6) and 2.074(6) Å are comparably long and indicate mainly σ‐bonding between the NHC ligand and the electron deficient d2 metal atom. The chloro ligand in [ReNCl(HLPh)(LPh)] is labile and can be replaced by ligands such as pseudohalides or monoanionic thiolates such as diphenyldithiophosphinate (Ph2PS2?) or pyridine‐2‐thiolate (pyS?). X‐ray structure analyses of [ReN(CN)(HLPh)(LPh)] and [ReN(pyS)(HLPh)(LPh)] show that the bonding situation of the NHC ligands (Re–C(carbene) distances between 2.086(3) and 2.130(3) Å) in the product is not significantly influenced by the ligand exchange. The potentially bidentate pyS? ligand is solely coordinated via its thiolato functionality. Hydrogen atoms of each one of the phenyl rings come close to the unoccupied sixth coordination positions of the rhenium atoms in the solid state structures of all complexes. Re–H distances between 2.620 and 2.712Å do not allow to discuss bonding, but with respect to the strong trans labilising influence of “N3?”, weak interactions are indicated.  相似文献   

20.
According to the covalent bond classification (CBC) method, two‐electron donors are defined as L‐type ligands, one‐electron donors as X‐type ligands, and two‐electron acceptors as Z‐type ligands. These three ligand functions are usually associated to the nature of the ligating atom, with phosphine, alkyl, and borane groups being prototypical examples of L‐, X‐ and Z‐ligands, respectively. A new SbNi platform is reported in which the ligating Sb atom can assume all three CBC ligand functions. Using both experimental and computational data, it is shown that PhICl2 oxidation of (o‐(Ph2P)C6H4)3SbNi(PPh3) ( 1 ) into [(o‐(Ph2P)C6H4)3ClSb]NiCl ( 2 ) is accompanied by a conversion of the stibine L‐type ligand of 1 into a stiboranyl X‐type ligand in 2 . Furthermore, the reaction of 2 with the catecholate dianion in the presence of cyclohexyl isocyanide results in the formation of [(o‐(Ph2P)C6H4)3(o‐O2C6H4Sb)]Ni(CNCy) ( 4 ), a complex featuring a nickel atom coordinated by a Lewis acidic, Z‐type, stiborane ligand.  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