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.
Iridium(I) and Iridium(III) Complexes with Triisopropylarsane as Ligand The ethene complex trans‐[IrCl(C2H4)(AsiPr3)2] ( 2 ), which was prepared from [IrCl(C2H4)2]2 and AsiPr3, reacted with CO and Ph2CN2 by displacement of ethene to yield the substitution products trans‐[IrCl(L)(AsiPr3)2] ( 3 : L = CO; 4 : L = N2). UV irradiation of 2 in the presence of acetonitrile gave via intramolecular oxidative addition the hydrido(vinyl)iridium(III) compound [IrHCl(CH=CH2)(CH3CN)(AsiPr3)2] ( 5 ). The reaction of 2 with dihydrogen led under argon to the formation of the octahedral complex [IrH2Cl(C2H4)(AsiPr3)2] ( 7 ), whereas from 2 under 1 bar H2 the ethene‐free compound [IrH2Cl(AsiPr3)2] ( 6 ) was generated. Complex 6 reacted with ethene to afford 7 and with pyridine to give [IrH2Cl(py)(AsiPr3)2] ( 8 ). The mixed arsane(phosphane)iridium(I) compound [IrCl(C2H4)(PiPr3)(AsiPr3)] ( 11 ) was prepared either from the dinuclear complex [IrCl(C2H4)(PiPr3)]2 ( 9 ) and AsiPr3 or by ligand exchange from [IrCl(C2H4)(PiPr3)(SbiPr3)] ( 10 ) und triisopropylarsane. The molecular structure of 5 was determined by X‐ray crystallography.  相似文献   

2.
The Dihydridoiridium(III) Complex [IrH2Cl(P i Pr3)2] as a Molecular Building Block for Unsymmetrical Binuclear Rhodium–Iridium and Iridium–Iridium Compounds The title compound [IrH2Cl(PiPr3)2] ( 3 ) reacts with the chloro‐bridged dimers [RhCl(PiPr3)2]2 ( 1 ) and [IrCl(C8H14)(PiPr3)]2 ( 5 ) by cleavage of the Cl‐bridges to give the unsymmetrical binuclear complexes 4 and 6 with Rh(μ‐Cl)2Ir and Ir(μ‐Cl)2Ir as the central building block. The reactions of 3 with the bis(cyclooctene) and (1,5‐cyclooctadiene) compounds [MCl(C8H14)2]2 ( 7 , 8 ) and [MCl(η4‐C8H12)]2 ( 9 , 10 ) (M = Rh, Ir) occur analogously and afford the rhodium(I)‐iridium(III) and iridium(I)‐iridium(III) complexes 11 – 14 in 70–80% yield. Treatment of [(η4‐C8H12)M(μ‐Cl)2IrH2(PiPr3)2] ( 13 , 14 ) with phenylacetylene leads to the formation of the substitution products [(η4‐C8H12)M(μ‐Cl)2IrH(C≡CPh)(PiPr3)2] ( 15 , 16 ) without changing the central molecular core. Similarly, the compound [(η4‐C8H12)Rh(μ‐Br)2IrH(C≡CPh)(PiPr3)2] ( 18 ) has been prepared; it was characterized by X‐ray crystallography.  相似文献   

3.
Synthesis and Dynamic Behaviour of [Rh2(μ-H)3H2(PiPr3)4]+. Contributions to the Reactivity of the Tetrahydridodirhodium Complex [Rh2H4(PiPr3)4] An improved synthesis of [Rh2H4(PiPr3)4] ( 2 ) from [Rh(η3-C3H5)(PiPr3)2] ( 1 ) or [Rh(η3-CH2C6H5)(PiPr3)2] ( 3 ) and H2 is described. Compound 2 reacts with CO or CH3OH to give trans-[RhH(CO)(PiPr3)2] ( 4 ) and with ethene/acetone to yield a mixture of 4 and trans-[RhCH3(CO)(PiPr3)2] ( 5 ). The carbonyl(methyl) complex 5 has also been prepared from trans-[RhCl(CO)(PiPr3)2] ( 6 ) and CH3MgI. Whereas the reaction of 2 with two parts of CF3CO2H leads to [RhH22-O2CCF3) · (PiPr3)2] ( 8 ), treatment of 2 with one equivalent of CF3CO2H in presence of NH4PF6 gives the dinuclear compound [Rh2H5(PiPr3)4]PF6 ( 9a ). The reactions of 2 with HBF4 and [NO]BF4 afford the complexes [Rh2H5(PiPr3)4]BF4 ( 9b ) and trans-[RhF(NO)(PiPr3)2]BF4 ( 11 ), respectively. In solution, the cation [Rh2(μ-H)3H2(PiPr3)4]+ of the compounds 9a and 9b undergoes an intramolecular rearrangement in which the bridging hydrido and the phosphane ligands are involved.  相似文献   

4.
Synthesis, Structure, and Photochemical Behavior of Olefine Iridium(I) Complexes with Acetylacetonato Ligands The bis(ethene) complex [Ir(κ2‐acac)(C2H4)2] ( 1 ) reacts with tertiary phosphanes to give the monosubstitution products [Ir(κ2‐acac)(C2H4)(PR3)] ( 2 – 5 ). While 2 (R = iPr) is inert toward PiPr3, the reaction of 2 with diphenylacetylene affords the π‐alkyne complex [Ir(κ2‐acac)(C2Ph2)(PiPr3)] ( 6 ). Treatment of [IrCl(C2H4)4] with C‐functionalized acetylacetonates yields the compounds [Ir(κ2‐acacR1,2)(C2H4)2] ( 8 , 9 ), which react with PiPr3 to give [Ir(κ2‐acacR1,2)(C2H4)(PiPr3)] ( 10 , 11 ) by displacement of one ethene ligand. UV irradiation of 5 (PR3 = iPr2PCH2CO2Me) and 11 (R2 = (CH2)3CO2Me) leads, after addition of PiPr3, to the formation of the hydrido(vinyl)iridium(III) complexes 7 and 12 . The reaction of 2 with the ethene derivatives CH2=CHR (R = CN, OC(O)Me, C(O)Me) affords the compounds [Ir(κ2‐acac)(CH2=CHR)(PiPr3)] ( 13 – 15 ), which on photolysis in the presence of PiPr3 also undergo an intramolecular C–H activation. In contrast, the analogous complexes [Ir(κ2‐acac)(olefin)(PiPr3)] (olefin = (E)‐C2H2(CO2Me)2 16 , (Z)‐C2H2(CO2Me)2 17 ) are photochemically inert.  相似文献   

5.
The hydro­lysis product [Ga2(C3H7)4(OH)2]·C14H32N4, derived from the tetrakis­(triiso­propyl­gallium)–1,4,8,11‐tetra­methyl‐1,4,8,11‐tetra­aza­cyclo­tetra­decane (1/1) adduct, consists of a centrosymmetric [iPr2Ga(μ‐OH)]2 unit hydrogen bonded through the hydroxyl group to a nitro­gen on an adjacent centrosymmetric 1,4,8,11‐tetra­methyl‐1,4,8,11‐tetra­aza­cyclo­tetra­decane molecule, resulting in the generation of a molecular chain through the crystal.  相似文献   

6.
The two title compounds, [Mo2Ir2(C6H7)2(CO)10] and [Mo2Ir2(C9H13)2(CO)10]·0.5CH2Cl2, respectively, or collectively [Mo2Ir2(μ‐CO)3(CO)75‐C5H5?nMen)2] (n = 1 or 4), have a pseudo‐tetrahedral Mo2Ir2 core geometry, an η5‐­C5H5?nMen group ligating each Mo atom, bridging carbonyls spanning the edges of an MoIr2 face and seven terminally bound carbonyl groups.  相似文献   

7.
The reactions of K[(2,6‐iPr2C6H3‐O)2POO] either with LaCl3(H2O)7 or with Nd(NO3)3(H2O)6 in a 3:1 molar ratio, followed by vacuum drying and recrystallization from alkanes, have led to the formation of diaquapentakis[bis(2,6‐diisopropylphenyl) phosphato]‐μ‐hydroxido‐dilanthanum hexane disolvate, [La2(C24H34O4P)5(OH)(H2O)2]·2C6H14, ( 1 )·2(hexane), and tetraaquatetrakis[bis(2,6‐diisopropylphenyl) phosphato]‐μ‐hydroxido‐dineodymium bis(2,6‐diisopropylphenyl) phosphate heptane disolvate, [Nd2(C24H34O4P)4(OH)(H2O)4]·2C6H14, ( 2 )·2(heptane). The compounds crystalize in the P21/n and P space groups, respectively. The diaryl‐substituted organophosphate ligand exhibits three different coordination modes, viz. κ2O,O′‐terminal [in ( 1 ) and ( 2 )], κO‐terminal [in ( 1 )] and μ2‐κ1O1O′‐bridging [in ( 1 ) and ( 2 )]. Binuclear structures ( 1 ) and ( 2 ) are similar and have the same unique Ln2(μ‐OH)(μ‐OPO)2 core. The structure of ( 2 ) consists of an [Nd2{(2,6‐iPr2C6H3‐O)2POO}4(OH)(H2O)4]+ cation and a [(2,6‐iPr2C6H3‐O)2POO] anion, which are bound via four intermolecular O—H…O hydrogen bonds. The molecular structure of ( 1 ) displays two O—H…O hydrogen bonds between OH/H2O ligands and a κ1O‐terminal organophosphate ligand, which resembles, to some extent, the `free' [(2,6‐iPr2C6H3‐O)2POO] anion in ( 2 ). NMR studies have shown that the formation of ( 1 ) undoubtedly occurs due to intramolecular hydrolysis during vacuum drying of the aqueous La tris(phosphate) complex. Catalytic experiments have demonstrated that the presence of the coordinated hydroxide anion and water molecules in precatalyst ( 2 ) substantially lowered the catalytic activity of the system prepared from ( 2 ) in butadiene and isoprene polymerization compared to the catalytic system based on the neodymium tris[bis(2,6‐diisopropylphenyl) phosphate] complex, which contains neither OH nor H2O ligands.  相似文献   

8.
Mono‐ and Dinuclear Rhodium Complexes with Arsino(phosphino)methanes in Different Coordination Modes The cyclooctadiene complex [Rh(η4‐C8H12)(κ2tBu2AsCH2PiPr2)](PF6) ( 1a ) reacts with CO and CNtBu to give the substitution products [Rh(L)22tBu2AsCH2PiPr2)](PF6) ( 2 , 3 ). From 1a and Na(acac) in the presence of CO the neutral compound [Rh(κ2‐acac)(CO)(κ‐PtBu2AsCH2PiPr2)] ( 4 ) is formed. The reactions of 1a , the corresponding B(ArF)4‐salt 1b and [Rh(η4‐C8H12)(κ2iPr2AsCH2PiPr2)](PF6) ( 5 ) with acetonitrile under a H2 atmosphere affords the complexes [Rh(CH3CN)22‐R2AsCH2PiPr2)]X ( 6a , 6b , 7 ), of which 6a (R = tBu; X = PF6) gives upon treatment with Na(acac‐f6) the bis(chelate) compound [Rh(κ2‐acac‐f6)(κ2tBu2AsCH2PiPr2)] ( 8 ). From 8 and CH3I a mixture of two stereoisomers of composition [Rh(CH3)I(κ2‐acac‐f6)(κ2tBu2AsCH2PiPr2)] ( 9/10 ) is generated by oxidative addition, and the molecular structure of the racemate 9 has been determined. The reactions of 1a and 5 with CO in the presence of NaCl leads to the formation of the “A‐frame” complexes [Rh2(CO)2(μ‐Cl)(μ‐R2AsCH2PiPr2)2](PF6) ( 11 , 12 ), which have been characterized crystallographically. From 11 and 12 the dinuclear substitution products [Rh2(CO)2(μ‐X)(μ‐R2AsCH2PiPr2)2](PF6) ( 13 ‐ 16 ) are obtained by replacing the bridging chloride for bromide, hydride or hydroxide, respectively. While 12 (R = iPr) reacts with NaI to give the related “A‐frame” complex 18 , treatment of 11 (R = tBu) with NaI yields the mononuclear chelate compound [RhI(CO)(κ2tBu2AsCH2PiPr2)] ( 20 ). The reaction of 20 with CH3I affords the acetyl complex [RhI2{C(O)CH3}(κ2tBu2AsCH2PiPr2)] ( 21 ) with five‐coordinate rhodium atom.  相似文献   

9.
Treating [Cp*V(μ‐Cl)2]3 (Cp* = C5Me5) and [(2,6‐i‐Pr2C6H3N)2MoMe2], respectively, with Me3SnF afforded the title compounds [Cp*V(μ‐F)2]4 ( 1 ) and [(2,6‐i‐Pr2C6H3N)2MoF2] · THF ( 2 ). 1 has a tetrameric structure, in which four V atoms can be regarded as being arranged at the vertices of a distorted tetrahedron, with four long edges bridged by one F atom and each of the other two short edges bridged by two F atoms with a mean V–F bond length of 2.00 Å. A hydrolyzed product of 2 , [(2,6‐i‐Pr2C6H3N)6Mo43‐F)2Me2(μ‐O)4] ( 3 ) was characterized by elemental analyses and X‐ray single crystal study. The X‐ray diffraction analysis reveals that 3 has a unique tetranuclear structure, containing two five and two six coordinated Mo atoms connecting each other by four μ‐O and two μ3‐F atoms. The geometries around the two Mo atoms can be described having distorted trigonal bipyramidal and distorted octahedral coordination spheres, respectively. The Mo–(μ‐O) bond lengths are 1.813 Å (average) for five coordinated Mo atoms and 2.030 Å (average) for those of six coordinated, respectively, indicating an additional π bonding between five coordinated Mo atoms and the μ‐O atoms. The Mo–(μ3‐F) distances range from 2.291 to 2.352 Å.  相似文献   

10.
The aromatic osmacyclopropenefuran bicycles [OsTp{κ3‐C1,C2,O‐(C1H2C2CHC(OEt)O)}(PiPr3)]BF4 (Tp=hydridotris(1‐pyrazolyl)borate) and [OsH{κ3‐C1,C2,O‐(C1H2C2CHC(OEt)O)}(CO)(PiPr3)2]BF4, with the metal fragment in a common vertex between the fused three‐ and five‐membered rings, have been prepared via the π‐allene intermediates [OsTp(η2‐CH2=CCHCO2Et)(OCMe2)(PiPr3)]BF4 and [OsH(η2‐CH2=CCHCO2Et)(CO)(OH2)(PiPr3)2]BF4, and their aromaticity analyzed by DFT calculations. The bicycle containing the [OsH(CO)(PiPr3)2]+ metal fragment is a key intermediate in the [OsH(CO)(OH2)2(PiPr3)2]BF4‐catalyzed regioselective anti‐Markovnikov hydration of ethyl buta‐2,3‐dienoate to ethyl 4‐hydroxycrotonate.  相似文献   

11.
In the crystals of bis(pyridine‐N)tetrakis(μ‐trimethylsilylacetato‐O:O′)dicopper(II), [Cu2(C5H11O2Si)4(C5H5N)2], (I), the dinuclear CuII complexes have cage structures with Cu?Cu distances of 2.632 (1) and 2.635 (1) Å. In the crystals of bis(2‐­methylpyridine‐N)tetrakis(μ‐trimethylsilylacetato‐O:O′)dicopper(II), [Cu2(C5H11O2Si)4(C6H7N)2], (II), bis­(3‐methylpyridine‐N)tetrakis(μ‐trimethylsilylacetato‐O:O′)dicopper(II), [Cu2(C5H11O2Si)4(C6H7N)2], (III), and bis(quinoline‐N)­tetrakis(μ‐­trimethylsilylacetato‐O:O′)dicopper(II), [Cu2(C5H11O2Si)4(C9H7N)2], (IV), the centrosymmetric dinuclear CuII complexes have a cage structure with Cu?Cu distances of 2.664 (1), 2.638 (3) and 2.665 (1) Å, respectively. In the crystals of catena‐poly­[tetrakis(μ‐trimethylsilylacetato‐O:O′)dicopper(II)], [Cu2(C5H11O2Si)4]n, (V), the dinuclear CuII units of a cage structure are linked by the cyclic Cu—O bonds at the apical positions to form a linear chain by use of a glide translation.  相似文献   

12.
The title compound, [Zn3(C9H21SiS)6] or [(iPr3SiS)Zn(μ‐SSiiPr3)2Zn(μ‐SSiiPr3)2Zn(SSiiPr3)], is the first structurally characterized homoleptic silanethiolate complex of zinc. A near‐linear arrangement of three ZnII ions is observed, the metals at the ends being three‐coordinate with one terminally bound silanethiolate ligand. The central ZnII ion is four‐coordinate and tetrahedral, with two bridging silanethiolate ligands joining it to each of the two peripheral ZnII ions. The nonbonding intermetallic distances are 3.1344 (11) and 3.2288 (12) Å, while the Zn...Zn...Zn angle is 172.34 (2)°. A trimetallic silanethiolate species of this type has not been previously identified by X‐ray crystallography for any element.  相似文献   

13.
Crystals of mononuclear tris[bis(2,6‐diisopropylphenyl) phosphato‐κO]pentakis(methanol‐κO)lanthanide methanol monosolvates of lanthanum, [La(C24H34O4P)3(CH3OH)5]·CH3OH, ( 1 ), cerium, [Ce(C24H34O4P)3(CH3OH)5]·CH3OH, ( 2 ), and neodymium, [Nd(C24H34O4P)3(CH3OH)5]·CH3OH, ( 3 ), have been obtained by reactions between LnCl3(H2O)n (n = 6 or 7) and lithium bis(2,6‐diisopropylphenyl) phosphate in a 1:3 molar ratio in methanol media. Compounds ( 1 )–( 3 ) crystallize in the monoclinic P21/c space group and have isomorphous crystal structures. All three bis(2,6‐diisopropylphenyl) phosphate ligands display a κO‐monodentate coordination mode. The coordination number of the metal atom is 8. Each [Ln{O2P(O‐2,6‐iPr2C6H3)2}3(CH3OH)5] molecular unit exhibits four intramolecular O—H…O hydrogen bonds, forming six‐membered rings. The unit forms two intermolecular O—H…O hydrogen bonds with one noncoordinating methanol molecule. All six hydroxy H atoms are involved in hydrogen bonding within the [Ln{O2P(O‐2,6‐iPr2C6H3)2}3(CH3OH)5]·CH3OH unit. This, along with the high steric hindrance induced by the three bulky diaryl phosphate ligands, prevents the formation of a hydrogen‐bond network. Complexes ( 1 )–( 3 ) exhibit disorder of two of the isopropyl groups of the phosphate ligands. The cerium compound ( 2 ) demonstrates an essential catalytic inhibition in the thermal decomposition of polydimethylsiloxane in air at 573 K. Catalytic systems based on the neodymium complex tris[bis(2,6‐diisopropylphenyl) phosphato‐κO]neodymium, ( 3′ ), which was obtained as a dry powder of ( 3 ) upon removal of methanol, display a high catalytic activity in isoprene and butadiene polymerization.  相似文献   

14.
Reaction of potassium salt of N‐aryliminopyrrole ligand [2‐(2, 6‐iPr2C6H3N=CH)–C4H3NK] ( 1 ) with samarium tris‐boro‐hydride [Sm(BH4)3(THF)3] gave a samarium ate complex [η2‐{2‐(2, 6‐iPr2C6H3N=CH)–C4H3N}3Sm(η1‐BH4){K(THF)6] ( 2 ); whereas similar treatment with erbium borohydride [Er(BH4)3(THF)3] afforded the mono(iminopyrrolyl) complex [η2‐{2‐(2, 6‐iPr2C6H3N=CH)–C4H3N}Er(η3‐BH4)2(THF)2] ( 3 ). In the solid‐state structures, the samarium complex 2 shows a rarely observed η1 and the erbium complex 3 shows a usual η3 coordination mode of the borohydrido ligand.  相似文献   

15.
The reaction of [PtCl2(COD)] (COD=1,5-cyclooctadiene) with diisopropyl-2-(3-methyl)indolylphosphine (iPr2P(C9H8N)) led to the formation of the platinum(ii ) chlorido complexes, cis-[PtCl2{iPr2P(C9H8N)}2] ( 1 ) and trans-[PtCl2{iPr2P(C9H8N)}2] ( 2 ). The cis-complex 1 reacted with NEt3 yielding the complex cis-[PtCl{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}] ( 3 ) bearing a cyclometalated κ2-(P,N)-phosphine ligand, while the isomer 2 with a trans-configuration did not show any reactivity towards NEt3. Treatment of 1 or 3 with (CH3)4NF (TMAF) resulted in the formation of the twofold cyclometalated complex cis-[Pt{κ2-(P,N)-iPr2P(C9H7N)}2] ( 4 ). The molecular structures of the complexes 1–4 were determined by single-crystal X-ray diffraction. The fluorido complex cis-[PtF{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}] ⋅ (HF)4 ( 5 ⋅ (HF)4) was formed when complex 4 was treated with different hydrogen fluoride sources. The Pt(ii ) fluorido complex 5 ⋅ (HF)4 exhibits intramolecular hydrogen bonding in its outer coordination sphere between the fluorido ligand and the NH group of the 3-methylindolyl moiety. In contrast to its chlorido analogue 3 , complex 5 ⋅ (HF)4 reacted with CO or the ynamide 1-(2-phenylethynyl)-2-pyrrolidinone to yield the complexes trans-[Pt(CO){κ2-(P,C)-iPr2P(C9H7NCO)}{iPr2P(C9H8N)}][F(HF)4] ( 7 ) and a complex, which we suggest to be cis-[Pt{C=C(Ph)OCN(C3H6)}{κ2-(P,N)-iPr2P(C9H7N)}{iPr2P(C9H8N)}][F(HF)4] ( 9 ), respectively. The structure of 9 was assigned on the basis of DFT calculations as well as NMR and IR data. Hydrogen bonding of HF and NH to fluoride was proven to be crucial for the existence of 7 and 9 .  相似文献   

16.
Strategies for the synthesis of highly electrophilic AuI complexes from either hydride‐ or chloride‐containing precursors have been investigated by employing sterically encumbered Dipp‐substituted expanded‐ring NHCs (Dipp=2,6‐iPr2C6H3). Thus, complexes of the type (NHC)AuH have been synthesised (for NHC=6‐Dipp or 7‐Dipp) and shown to feature significantly more electron‐rich hydrides than those based on ancillary imidazolylidene donors. This finding is consistent with the stronger σ‐donor character of these NHCs, and allows for protonation of the hydride ligand. Such chemistry leads to the loss of dihydrogen and to the trapping of the [(NHC)Au]+ fragment within a dinuclear gold cation containing a bridging hydride. Activation of the hydride ligand in (NHC)AuH by B(C6F5)3, by contrast, generates a species (at low temperatures) featuring a [HB(C6F5)3]? fragment with spectroscopic signatures similar to the “free” borate anion. Subsequent rearrangement involves B?C bond cleavage and aryl transfer to the carbophilic metal centre. Under halide abstraction conditions utilizing Na[BArf4] (Arf=C6H3(CF3)2‐3,5), systems of the type [(NHC)AuCl] (NHC=6‐Dipp or 7‐Dipp) generate dinuclear complexes [{(NHC)Au}2(μ‐Cl)]+ that are still electrophilic enough at gold to induce aryl abstraction from the [BArf4]? counterion.  相似文献   

17.
The two dinuclear IrI complexes [Ir2(μ‐Cl)2 {(R)‐(S)‐PPF‐PPh2}2] ( 1 ; (R)‐(S)‐PPF‐PPh2=(S)‐1‐(diphenylphosphino)‐2‐[(R)‐1‐(diphenylphosphino)ethyl]ferrocene and [Ir2(μ‐Cl)2{(R)‐binap}2] ( 3 ; (R)‐binap=(R)‐[1,1′‐binaphthalene]‐2,2′‐diylbis[diphenylphosphine]) smoothly react with 4 equiv. of the lithium salt of aniline to afford the new bis(anilido)iridate(I) (=bis(benzenaminato)iridate(1‐)) complexes Li[Ir(NHPh)2{(R)‐(S)‐PPF‐PPh2}] ( 4 ) and Li[Ir(NHPh)2{(R)‐binap}] ( 5 ), respectively. The anionic complexes 4 and 5 react upon protonolysis to give the dinuclear aminato‐bridged derivatives [Ir2(μ‐NHPh)2{(R)‐(S)‐PPF‐PPh2}2] ( 6 ) and [Ir2(μ‐NHPh)2{(R)‐binap}2] ( 7 ), which were characterized by X‐ray crystallography. None of the new complexes 4 – 7 shows catalytic activity in the hydroamination of olefins.  相似文献   

18.
The coordination chemistry of the 1,2‐BN‐cyclohexanes 2,2‐R2‐1,2‐B,N‐C4H10 (R2=HH, MeH, Me2) with Ir and Rh metal fragments has been studied. This led to the solution (NMR spectroscopy) and solid‐state (X‐ray diffraction) characterization of [Ir(PCy3)2(H)22η2‐H2BNR2C4H8)][BArF4] (NR2=NH2, NMeH) and [Rh(iPr2PCH2CH2CH2PiPr2)(η2η2‐H2BNR2C4H8)][BArF4] (NR2=NH2, NMeH, NMe2). For NR2=NH2 subsequent metal‐promoted, dehydrocoupling shows the eventual formation of the cyclic tricyclic borazine [BNC4H8]3, via amino‐borane and, tentatively characterized using DFT/GIAO chemical shift calculations, cycloborazane intermediates. For NR2=NMeH the final product is the cyclic amino‐borane HBNMeC4H8. The mechanism of dehydrogenation of 2,2‐H,Me‐1,2‐B,N‐C4H10 using the {Rh(iPr2PCH2CH2CH2PiPr2)}+ catalyst has been probed. Catalytic experiments indicate the rapid formation of a dimeric species, [Rh2(iPr2PCH2CH2CH2PiPr2)2H5][BArF4]. Using the initial rate method starting from this dimer, a first‐order relationship to [amine‐borane], but half‐order to [Rh] is established, which is suggested to be due to a rapid dimer–monomer equilibrium operating.  相似文献   

19.
The two‐step one‐pot oxidative decarbonylation of [Fe2(S2C2H4)(CO)4(PMe3)2] ( 1 ) with [FeCp2]PF6, followed by addition of phosphane ligands, led to a series of diferrous dithiolato carbonyls 2 – 6 , containing three or four phosphane ligands. In situ measurements indicate efficient formation of 1 2+ as the initial intermediate of the oxidation of 1 , even when a deficiency of the oxidant was employed. Subsequent addition of PR3 gave rise to [Fe2(S2C2H4)(μ‐CO)(CO)3(PMe3)3]2+ ( 2 ) and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)2(PR3)2]2+ (R=Me 3 , OMe 4 ) as principal products. One terminal CO ligand in these complexes was readily substituted by MeCN, and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)3(MeCN)]2+ ( 5 ) and [Fe2(S2C2H4)(μ‐CO)(CO)(PMe3)4(MeCN)]2+ ( 6 ) were fully characterized. Relevant to the Hred state of the active site of Fe‐only hydrogenases, the unsymmetrical derivatives 5 and 6 feature a semibridging CO ligand trans to a labile coordination site.  相似文献   

20.
Reaction of carbene‐stabilized disilicon ( 1 ) with Fe(CO)5 gives the 1:1 adduct L:Si?Si[Fe(CO)4]:L (L:=C{N(2,6‐Pri2C6H3)CH}2) ( 2 ) at room temperature. At raised temperature, however, 2 may react with another equivalent of Fe(CO)5 to give L:Si[μ‐Fe2(CO)6](μ‐CO)Si:L ( 3 ) through insertion of both CO and Fe2(CO)6 into the Si2 core, which represents the first experimental realization of transition metal‐carbonyl‐mediated cleavage of a Si?Si double bond. The structures and bonding of both 2 and 3 have been investigated by spectroscopic, crystallographic, and computational methods.  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