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 312 毫秒
1.
The reaction of [ReBr(CO)5] with phosphite and phosphonite ligands in toluene yielded cis, mer‐[ReBr(CO)2L3] ( 2 : L = P(OMe)3 2a : P(OEt)3 2b : PPh(OMe)2 2c : PPh(OEt)2 2d ). Compounds 2c and 2d were also obtained, as were the phosphinite complexes 2e [L = PPh2(OMe)] and 2f [L = PPh2(OEt)], by reaction of the corresponding phosphorus ligand with trans, mer‐[ReBr(CO)3L2]. Compounds 2 were all characterized by elemental analysis, mass spectrometry and NMR spectroscopy, and the structures of 2b , 2c and 2d were determined by X‐ray diffractometry. Compounds 2a‐d are stable in chloroform and dichloromethane, but 2e and 2f are transformed into the corresponding trans, mer‐[ReBr(CO)3L2] complexes by a reaction for which a partial mechanism is put forward.  相似文献   

2.
The manganese carbonyl complex [MnBr(CO)3 L ] ( 1 ), where L = Ph2POCH2CH2OPPh2, was prepared by reacting [MnBr(CO)5] with the bidentate ligand 1, 2‐Bis(diphenylphosphinite)ethane. From this compound and the appropriate phosphite, phosphinite or phosphonite ligands were synthesized the complexes [MnBr(CO)2 LL ′], where L ′ = P(OMe)3 ( 2 ) or P(OEt)3 ( 3 ) and [MnBr(CO)3 L ′2], where L ′ =PPh(OEt)2 ( 4 ) or PPh2(OEt) ( 5 ). The obtained compounds have been characterized by elemental analysis, mass spectrometry, IR and NMR (1H, 13C and 31P) spectroscopies and X‐ray diffractometry for the complexes 1 , 4 and 5 .  相似文献   

3.
Ethylene complexes [OsH(η2‐CH2=CH2)L4]Y ( 1 , 2 ) [L = PPh(OEt)2, P(OEt)3; Y = OTf, BPh4] were prepared by reacting the dihydride OsH2L4 first with methyl triflate CH3OTf and then with ethylene (1 atm). Alternatively, the compound [OsH(η2‐CH2=CH2){PPh(OEt)2}4]OTf was prepared by allowing the dinitrogen derivative [OsH(N2){PPh(OEt)2}4]OTf to react with ethylene. Acrylonitrile CH2=C(H)CN reacts with OsH(OTf)L4 [L = P(OEt)3] to give the complex [OsH{κ1‐NCC(H)=CH2}{P(OEt)3}4]BPh4 ( 3 ). The complexes were characterized spectroscopically (IR and 1H, 13C, 31P NMR) and by X‐ray crystal structure determination of the [OsH(η2‐CH2=CH2){PPh(OEt)2}4]BPh4 derivative.  相似文献   

4.
The mono- and binuclear aryldiazene complexes [Re(C6H5N=NH)(CO)5-nPn]BY4 (1-5) and [(Re(CO)5-nPn)2-(mu-HN=NAr-ArN=NH)](BY4)2 (6-12) [P = P(OEt)3, PPh(OEt)2, PPh2OEt; n = 1-4; Ar-Ar = 4,4'-C6H4-C6H4, 4,4'-(2-CH3)C6H3-C6H3(2-CH3), 4,4'-C6H4-CH2-C6H4; Y = F, Ph) were prepared by reacting the hydride species ReH(CO)5-nPn with the appropriate mono- and bis(aryldiazonium) cations. These compounds, as well as other prepared compounds, were characterized spectroscopically (IR; 1H, 31P, 13C, and 15N NMR data), and 1a was also characterized by an X-ray crystal structure determination. [Re(C6H5N=NH)(CO)(P(OEt)3)4]BPh4 (1a) crystallizes in space group P1 with a = 15.380(5) A, b = 13.037(5) A, c = 16.649(5) A, alpha = 90.33(5) degrees, beta = 91.2(1) degrees, gamma = 89.71(9) degrees, and Z = 2. The "diazene-diazonium" complexes [M(CO)3P2(HN=NAr-ArN identical to N)](BF4)2 (13-15, 17) [M = Re, Mn; P = PPh2OEt, PPh2OMe, PPh3; Ar-Ar = 4,4'-C6H4-C6H4, 4,4'-C6H4-CH2-C6H4] and [Re(CO)4(PPh2OEt)(4,4'-HN=NC6H4-C6H4N identical to N)](BF4)2 (16b) were synthesized by allowing the hydrides MH(CO)3P2 or ReH(CO)4P to react with equimolar amounts of bis(aryldiazonium) cations under appropriate conditions. Reactions of diazene-diazonium complexes 13-17 with the metal hydrides M2H2P'4 and M2'H(CO)5-nP"n afforded the heterobinuclear bis(aryldiazene) derivatives [M1(CO)3P2(mu-HN=NAr-ArN=NH)M2HP'4](BPh4)2 (ReFe, ReRu, ReOs, MnRu, MnOs) and [M1(CO)3P2(mu-HN=NAr-ArN=NH)M2'(CO)5-nP"n](BPh4)2 (ReMn, MnRe) [M1 = Re, Mn; M2 = Fe, Ru, Os; M2' = Mn, Re; P = PPh2OEt, PPh2OMe; P',P" = P(OEt)3, PPh(OEt)2; Ar-Ar = 4,4'-C6H4-C6H4, 4,4'-C6H4-CH2-C6H4; n = 1, 2]. The heterotrinuclear complexes [Re(CO)3(PPh2OEt)2(mu-4,4'-HN=NC6H4-C6H4N=NH)M(P(OEt)3)4(mu-4,4'-HN=NC6H4- C6H4N=NH)Mn(CO)3(PPh2OEt)2](BPh4)4 (M = Ru, Os) (ReRuMn, ReOsMn) were obtained by reacting the heterobinuclear complexes ReRu and ReOs with the appropriate diazene-diazonium cations. The heterobinuclear complex with a bis(aryldiazenido) bridging ligand [Mn(CO)2(PPh2OEt)2(mu-4,4'-N2C6H4-C6H4N2)Fe(P(OEt)3)4]BPh4 (MnFe) was prepared by deprotonating the bis(aryldiazene) compound [Mn(CO)3(PPh2OEt)2(mu-4,4'-HN=NC6H4-C6H4N=NH)Fe(4- CH3C6H4CN)(P(OEt)3)4](BPh4)3. Finally, the binuclear compound [Re(CO)3(PPh2OEt)2(mu-4,4'-HN=NC6H4-C6H4N2)Fe(CO)2(P(OPh)3)2](BPh4)2 (ReFe) containing a diazene-diazenido bridging ligand was prepared by reacting [Re(CO)3(PPh2OEt)2(4,4'-HN=NC6H4-C6H4N identical to N)]+ with the FeH2(CO)2(P(OPh)3)2 hydride derivative. The electrochemical reduction of mono- and binuclear aryldiazene complexes of both rhenium (1-12) and the manganese, as well as heterobinuclear ReRu and MnRu complexes, was studied by means of cyclic voltammetry and digital simulation techniques. The electrochemical oxidation of the mono- and binuclear aryldiazenido compounds Mn(C6H5N2)(CO)2P2 and (Mn(CO)2P2)2(mu-4,4'-N2C6H4-C6H4N2) (P = PPh2OEt) was also examined. Electrochemical data show that, for binuclear compounds, the diazene bridging unit allows delocalization of electrons between the two different redox centers of the same molecule, whereas the two metal centers behave independently in the presence of the diazenido bridging unit.  相似文献   

5.
The pseudo‐square‐planar complexes [Rh(cod)(Hbbtm)]BF4 ( 3 ), [Rh(bbte)(cod)]BF4 ( 4 ), [Rh(CO)2(Hbbtm)]BF4 ( 5 ), [Rh(bbte)(CO)2]BF4 ( 6 ), [Rh(bbtm)(cod)] ( 7 ) and [Rh(bbtm)(CO)2] ( 8 ) (Hbbtm=bis(benzothiazol‐2‐yl)methane=2,2′‐methylenebis[benzothiazole], bbte=bis(benzothiazol‐2‐yl)ethane=2,2′‐(ethane‐1,2‐diyl)bis[benzothiazole], and cod=cycloocta‐1,5‐diene) were synthesized and characterized. Diastereotopic protons were observed for the protons at the bridge in the 1H‐NMR of 3 and 5 . Twisting of the ethane‐1,2‐diyl bridge in 4 and 6 effects chemical equivalence of the CH2 groups in solution. Unusually large downfield shifts occur on coordination of the deprotonated ligand Hbbtm as the negative charge is delocalized in 7 and 8 . The NMR signals of the cod ligand in 4 could be differentiated. The X‐ray crystal structures of 3, 4 , and 6 are reported.  相似文献   

6.
Multifaceted Coordination Chemistry of Vanadium(V): Substitution, Rearrangement Reactions, and Condensation Reactions of Oxovanadium(V) Complexes of the Tripodal Oxygen Ligand LOMe? = [η5‐(C5H5)Co{P(OMe)2(O)}3]? The octahedral oxovanadium(V) complex [V(O)F2LOMe] of the tripodal oxygen ligand LOMe? = [η5‐(C5H5)Co{P(OMe)2(O)}3]? reacts with alcohols and phenol with substitution of one fluoride ligand to form alkoxo complexes [V(O)F(OR)LOMe], R = Me, Et, i‐Prop, Ph. In the presence of water, however, both fluoride ions are substituted and a complex with the composition VO2LOMe can be isolated. The crystal structure shows that the oxo‐bridged trimer [{V(O)(LOMe)O}3] was synthesized. In the presence of BF3 the fluoride ligand in the alkoxo‐complex [V(O)F(OEt)LOMe] can be exchanged for pyridine to yield [V(O)(OEt)pyLOMe]BF4. Analogous attempts to exchange the fluoride ligand for tetrahydrofuran and acetonitrile induces a rearrangement reaction that leads to the vanadium complex [V(O)(LOMe)2]BF4. The crystal structure of this compound has been determined. Its 1H and 31P‐NMR spectra show that it is a highly fluxional vanadium complex at ambient temperature in solution. The two tripodal ligands LOMe? coordinate the vanadium centre as bidentate or tridentate ligands. The exchange bidentate/tridentate becomes slow on the NMR time scale below about 200 K.  相似文献   

7.
Photoirradiation of a toluene solution of [ReH(CO)3(L)] [S. Bolaño, J. Bravo, R. Carballo, S. García-Fontán, U. Abram, E.M. Vázquez-López, Polyhedron 18 (1999) 1431-1436] [L = 1,2-bis(diphenylphosphinoxy)ethane] in the presence of PPhn(OR)3−n (n = 0, 1; R = Me, Et) leads to the replacement of a CO ligand by the corresponding monodentate phosphite or phosphonite ligand to give new hydride compounds of formula [ReH(CO)2(L)(L′)] [L′ = P(OMe)3 (1); P(OEt)3 (2); PPh(OMe)2 (3); PPh(OEt)2 (4)]. Protonation of compounds 1-4 in CD2Cl2, with HBF4.OMe2 or with HOOCCF3 at 193 K in a NMR tube, gave the corresponding dihydrogen complexes. When the temperature was increased from 193 to 293 K, the η2-H2 ligand was replaced by OMe2 or OOCCF3 groups (depending on the acid employed) to give new stable complexes and the loss of H2 gas.  相似文献   

8.
The NiII‐mediated tautomerization of the N‐heterocyclic hydrosilylcarbene L2Si(H)(CH2)NHC 1 , where L2=CH(C?CH2)(CMe)(NAr)2, Ar=2,6‐iPr2C6H3; NHC=3,4,5‐trimethylimidazol‐2‐yliden‐6‐yl, leads to the first N‐heterocyclic silylene (NHSi)–carbene (NHC) chelate ligand in the dibromo nickel(II) complex [L1Si:(CH2)(NHC)NiBr2] 2 (L1=CH(MeC?NAr)2). Reduction of 2 with KC8 in the presence of PMe3 as an auxiliary ligand afforded, depending on the reaction time, the N‐heterocyclic silyl–NHC bromo NiII complex [L2Si(CH2)NHCNiBr(PMe3)] 3 and the unique Ni0 complex [η2(Si‐H){L2Si(H)(CH2)NHC}Ni(PMe3)2] 4 featuring an agostic Si? H→Ni bonding interaction. When 1,2‐bis(dimethylphosphino)ethane (DMPE) was employed as an exogenous ligand, the first NHSi–NHC chelate‐ligand‐stabilized Ni0 complex [L1Si:(CH2)NHCNi(dmpe)] 5 could be isolated. Moreover, the dicarbonyl Ni0 complex 6 , [L1Si:(CH2)NHCNi(CO)2], is easily accessible by the reduction of 2 with K(BHEt3) under a CO atmosphere. The complexes were spectroscopically and structurally characterized. Furthermore, complex 2 can serve as an efficient precatalyst for Kumada–Corriu‐type cross‐coupling reactions.  相似文献   

9.
Reaction of bisalkylidyne cluster compounds [Fe3(CO)93‐CR)2] ( 1a—d ) ( a , R = H; b , R = F; c , R = Cl; d , R = Br) with the phosphaalkyne t‐C4H9‐C≡P ( 2 ) yield a single isomer of the phosphaferrole cluster [Fe3(CO)8][CR‐C(t‐Bu)‐P‐CR] ( 3a—d ). However, the three isomeric compounds [Fe3(CO)8][C(OEt)‐C(t‐Bu)‐P‐C(Me)] ( 5a ), [Fe3(CO)8][C(Me)‐C(t‐Bu)‐P‐C(OEt)] ( 5b ), and [Fe3(CO)8][C(OEt)‐C(Me)‐C(t‐Bu)‐P] ( 5c ) are obtained in the reaction of [Fe3(CO)93‐CMe)(μ3‐C‐OEt)] ( 4 ) with 2 . As the phosphaferroles 3 possess a lone pair of electrons at the phosphorus atom they can act as ligands. [Fe3(CO)8][CF‐C(t‐Bu)‐P‐CF]MLn ( 7a—c ) ( a , MLn = Cr(CO)5; b , MLn = CpMn(CO)2; c , MLn = Cp*Mn(CO)2) were formed from 3b and LnM(η2‐C8H14) ( 6a—c ). The dinuclear cluster [Fe2(CO)6][CF‐CF‐C(t‐Bu)‐PH(OMe)] ( 8 ) was obtained from 3b and NiCl2·6H2O in methanol. The structures of 3a—d , 5a—c , 7b , and 8 have been elucidated by X‐ray crystal structure determinations.  相似文献   

10.
This article deals with isomeric ruthenium complexes [RuIII(LR)2(acac)] (S=1/2) involving unsymmetric β‐ketoiminates (AcNac) (LR=R‐AcNac, R=H ( 1 ), Cl ( 2 ), OMe ( 3 ); acac=acetylacetonate) [R=para‐substituents (H, Cl, OMe) of N‐bearing aryl group]. The isomeric identities of the complexes, cct (ciscis‐trans, blue, a ), ctc (cis‐trans‐cis, green, b ) and ccc (ciscis‐cis, pink, c ) with respect to oxygen (acac), oxygen (L) and nitrogen (L) donors, respectively, were authenticated by their single‐crystal X‐ray structures and spectroscopic/electrochemical features. One‐electron reversible oxidation and reduction processes of 1 – 3 led to the electronic formulations of [RuIII(L)(L ? )(acac)]+ and [RuII(L)2(acac)]? for 1 +‐ 3 + (S=1) and 1? – 3? (S=0), respectively. The triplet state of 1 +‐ 3 + was corroborated by its forbidden weak half‐field signal near g≈4.0 at 4 K, revealing the non‐innocent feature of L. Interestingly, among the three isomeric forms ( a – c in 1 – 3 ), the ctc ( b in 2 b or 3 b ) isomer selectively underwent oxidative functionalization at the central β‐carbon (C?H→C=O) of one of the L ligands in air, leading to the formation of diamagnetic [RuII(L)(L ′ )(acac)] (L ′ =diketoimine) in 4 / 4′ . Mechanistic aspects of the oxygenation process of AcNac in 2 b were also explored via kinetic and theoretical studies.  相似文献   

11.
The catecholase activity of the dicopper(II) complexes [Cu2(L1)(μ‐OCH3)(NCCH3)2](PF6)2·H2O·CH3CN ( 1 ), [Cu2(L2)(μ‐OH)(MeOH)(NCCH3)](BF4)2 ( 2 ), [Cu2(L3)(μ‐OMe)(NCCH3)2](BF4)2·2CH3CN·H2O ( 3 ), [Cu2(L2)(μ‐OAc)2]BF4·H2O ( 4 ), [Cu2(L4)(μ‐OAc)2]ClO4 ( 5 ) and [Cu2(L5)(μ‐OMe)(NCCH3)3(OH2)](ClO4)2·2CH3OH·CH3CN ( 6 ) consisting of varying para‐substituted phenol ligands HL1 = 4‐trifluoromethyl‐2,6‐bis((4‐methylpiperazin‐1‐yl)methyl)phenol, HL2 = 4‐bromo‐2,6‐bis((4‐methyl‐1,4‐diazepan‐1‐yl)methyl)phenol, HL3 = 4‐bromo‐2‐((4‐methyl‐1,4‐diazepan‐1‐yl)methyl)‐6‐((4‐methylpiperazin‐1‐yl)methyl)phenol, HL4 = 2,6‐bis((4‐methylpiperazin‐1‐yl)methyl)‐4‐nitrophenol and HL5 = 4‐tert‐butyl‐2,6‐bis((4‐methylpiperazin‐1‐yl)methyl)phenol was studied. The main difference within the six complexes lies in the individual copper–copper separation that is enforced by the chelating side arms of the phenolate ligand entity and more importantly in the exogenous bridging solvent, hydroxide, methanolate or acetate ions. The distance between the copper cores varies from 2.94Å in 1 to 3.29Å in 5 . The catalytic activity of the complexes 1 – 6 towards the oxidation of 3,5‐di‐tert‐butylcatechol was determined spectrophotometrically by monitoring the increase of the 3,5–di‐tert‐butylquinone characteristic absorption band at about 400 nm over time saturated with O2. The complexes are able to oxidize the substrate 3,5‐di‐tert‐butylcatechol to the corresponding o‐quinone with distinct catalytic activity (kcat between 92 h?1 and 189 h?1), with an order of decreasing activity 6 > 5 > 1 , 2 , 4 ≥ 3 . A kinetic treatment of the data based on the Michaelis‐Menten approach was applied. A correlation of the catecholase activities with the variation of the para‐ substituents as well as other effects resulting from the copper core distances is discussed. [Cu2(L5)(μ‐OMe)(NCCH3)3(OH)2](ClO4)2·2CH3OH·CH3CN ( 6 ) exhibited the highest activity of the six complexes as a result of its high turnover rate.  相似文献   

12.
Treatment of bis(cyanamide) [M(N≡CNEt2)2L4](BPh4)2 and bis(cyanoguanidine) [M{N≡CN(H)C(NH2)=NH}2L4](BPh4)2 complexes [M = Fe, Ru, Os; L = P(OEt)3] with an excess of amine RNH2 (R = nPr, iPr) affords mixed‐ligand complexes with cyanamide and amine [M(NH2R)(N≡CNEt2)L4](BPh4)2 ( 1a – 5a ) and [M(NH2R){N≡CN(H)C(NH2)=NH}L4](BPh4)2 ( 1b , 2b ). The complexes were characterized by spectroscopy and X‐ray crystal structure determination of [M(NH2iPr)(N≡CNEt2){P(OEt)3}4](BPh4)2 [M = Ru ( 3a ), Os ( 5a )].  相似文献   

13.
Reactions of bis(phosphinimino)methanes H2C(PPh2NR)2 [R = SiMe3 (L1H), Ph (L2H), 2,6‐iPr2‐C6H3 (DIPP) (L3H)] with ZnR2 (R = Me, Et) yielded the corresponding bis(phosphinimino)methanide zinc complexes LZnMe [L2 ( 1 ), L3 ( 2 )] and LZnEt [L1 ( 3 ), L2 ( 4 ), and L3 ( 5 )]. Complexes 1 – 5 were characterized by heteronuclear NMR (1H, 13C, 31P) and IR spectroscopy, elemental analysis, and single‐crystal X‐ray diffraction.  相似文献   

14.
Reaction of a mixture of AgOAc, Lawesson's reagent [2,4‐bis(4‐methoxyphenyl)‐1,3‐dithiadiphosphetane‐2,4‐disulfide] and 1,3‐bis(diphenylphosphanyl)propane (dppp) under ultrasonic treatment gave the title compound, {[Ag(C9H12O2PS2)(C27H26P2)]·CHCl3}n, a novel one‐dimensional chain based on the in situ‐generated bipodal ligand [ArP(OEt)S2] (Ar = 4‐methoxyphenyl). The compound consists of bidentate bridging 1,3‐bis(diphenylphosphanyl)propane (dppp) and in situ‐generated bidentate chelating [ArP(OEt)S2] ligands. The dppp ligand links the [Ag{ArP(OEt)S2}] subunit to form an achiral one‐dimensional infinite chain. These achiral chains are packed into chiral crystals by virtue of van der Waals interactions. No π–π interactions are observed in the crystal structure.  相似文献   

15.
The cationic cluster complexes [Ru3(μ‐H)(μ‐κ2N,C‐L1 Me)(CO)10]+ ( 1 +; HL1 Me=N‐methylpyrazinium), [Ru3(μ‐H)(μ‐κ2N,C‐L2 Me)(CO)10]+ ( 2 +; HL2 Me=N‐methylquinoxalinium), and [Ru3(μ‐H)(μ‐κ2N,C‐L3 Me)(CO)10]+ ( 3 +; HL3 Me=N‐methyl‐1,5‐naphthyridinium), which contain cationic N‐heterocyclic ligands, undergo one‐electron reduction processes to become short lived, ligand‐centered, trinuclear, radical species ( 1 – 3 ) that end in the formation of an intermolecular C? C bond between the ligands of two such radicals, thus leading to neutral hexanuclear derivatives. These dimerization processes are selective, in the sense that they only occur through the exo face of the bridging ligands of trinuclear enantiomers of the same configuration, as they only afford hexanuclear dimers with rac structures (C2 symmetry). The following are the dimeric products that have been isolated by using cobaltocene as reducing agent: [Ru6(μ‐H)26‐κ4N2,C2‐(L1 Me)2}(CO)18] ( 5 ; from 1 +), [Ru6(μ‐H)26‐κ4N2,C2‐(L2 Me)2}(CO)18] ( 6 ; from 2 +), and [Ru6(μ‐H)24‐κ8N2,C6‐(L3 Me)2}(CO)18] ( 7 ; from 3 +). The structures of the final hexanuclear products depend on the N‐heterocyclic ligand attached to the starting materials. Thus, although both trinuclear subunits of 5 and 6 are face‐capped by their bridging ligands, the coordination mode of the ligand of 5 is different from that of the ligand of 6 . The trinuclear subunits of 7 are edge‐bridged by its bridging ligand. In the presence of moisture, the reduction of 3 + with cobaltocene also affords a trinuclear derivative, [Ru3(μ‐H)(μ‐κ2N,C‐L3′ Me)(CO)10] ( 8 ), whose bridging ligand (L3′ Me) results from the formal substitution of an oxygen atom for the hydrogen atom (as a proton) that in 3 + is attached to the C6 carbon atom of its heterocyclic ligand. The results have been rationalized with the help of electrochemical measurements and DFT calculations, which have also shed light on the nature of the odd‐electron species, 1 – 3 , and on the regioselectivity of their dimerization processes. It seems that the sort of coupling reactions described herein requires cationic complexes with ligand‐based LUMOs.  相似文献   

16.
Synthesis, crystal structure, thermal stability, and magnetic properties of mixed‐ligand complexes of cobalt(II) with ß‐diketonato (thd = C11H19O2?) and alkoxides (OR mainly OMe = methoxide = CH3O? or OEt = ethoxide = C2H5O?) are reported. Direct reaction between Co(thd)2 ( 1 ) and EtOH gives a new complex with the structural formula [Co4(thd)4(OEt)4] ( 2 ) whereas MeOH correspondingly reacts to [Co4(thd)4(OMe)4(MeOH)4] ( 3 ). The yield of these products decreases with increasing size of the R group owing to increased solubility of 1 in the alcohol. The structure of 2 is determined from single‐crystal X‐ray diffraction data. At 100 K 2 takes a monoclinic structure (space group C2/c): a = 15.108(2), b = 19.428(2), c = 21.240(3) Å, and β = 108.882(2)°. At 295 K 2 has transformed to a closely related orthorhombic structure (space group Fddd): a = 15.233(3), b = 19.712(3), c = 40.916(7) Å. Protracted hydrolysis accompanied by oxygenation of complexes 2 and 3 in laboratory air (viz. simultaneous exposure to moisture and oxygen) leads to a new complex 4 with empirical formula corresponding to [Co(thd)(OH)(O2)]. Magnetic susceptibility data show that Co takes the valence state II in all complexes 1 – 4 . For 4 this implies that dioxygen has to form an adduct‐like association to the rest of the complex. Unfortunately complex 4 has hitherto only been obtained in the amorphous state, but all here produced evidences point at 4 as a distinct entity and that products of 4 obtained from 2 and 3 are chemically identical (but differ somewhat with regard to short‐ and longer‐range order in the atomic arrangement). The interatomic distances in the crystal structure of complexes 1 – 3 are briefly discussed in terms of the bond‐valence concept.  相似文献   

17.
Syntheses of Oxovanadium(V) Halide Complexes Stabilized with Tripodal Oxygen Ligands LR = [η5‐(C5H5)Co{PR2(O)}3], R = OMe, OEt The sodium salts of the tripodal oxygen ligands LR = [η5‐(C5H5)Co{PR2(O)}3] (R = OMe, OEt) react with the oxovanadium halides V(O)F3 and V(O)Cl3 to yield deep red compounds of the type [V(O)X2LR]. Halide exchange reactions with [V(O)Cl2LOMe] und [V(O)F2LOMe] aiming at the preparation of the analogous bromide complex [V(O)Br2LOMe] led to the isomer [VO(LOMe)2][V(O)Br4]. The crystal structure of [V(O)Cl2LOMe] has been determined by single crystal x‐ray diffraction. The compound crystallizes in the monoclinic space group P21/n with a = 9.6332(8), b = 15.0312(11) and c = 15.3742(12)Å, β = 100.181(8)°. The coordination around vanadium is distorted octahedral.  相似文献   

18.
Synthesis, Structure, and Reactivity of η1‐ and η3‐Allyl Rhenium Carbonyls In (η3‐C3H5)Re(CO)4 one CO ligand can be substituted by PPh3, pyridine, isocyanide and benzonitrile. With 1,2‐bis(diphenylphosphino)ethylene, 1,1′‐bis(diphenylphosphino)ferrocene and 1,2‐bis(4‐pyridyl)ethane dinuclear ligand bridged complexes are obtained. The η3‐η1 conversion of the allyl ligand occurs on reaction of (η3‐C3H5)Re(CO)4 with the bidendate ligands 1,2‐bis(diphenylphosphino)ethane and 1,3‐bis(diphenylphosphino)propane and with 2,2′‐bipyridine (L–L) which gives the complexes (η1‐C3H5)Re(CO)3(L–L). By reaction of (η3‐C3H5)Re(CO)4 with bis(diphenylphosphino)methane the allyl group is protonated and under elemination of propene the complex (OC)3Re(Ph2PCHPPh2)(η1‐Ph2PCH2PPh2) ( 19 ) with a diphosphinomethanide ligand is formed. On heating solutions of (η3‐C3H5)Re(CO)4 and (η3‐C3H5)Re(CO)3(CN‐2,5‐Me2C6H3) ( 5 ) in methanol the methoxy bridged compounds Re4(CO)12(OH)(OMe)3 and Re2(CO)4(CN‐2,5‐Me2C6H3)4(μ‐OMe)2 ( 20 ) were isolated. The crystal structures of (η3‐C3H5)Re(CO)3(CNCH2SiMe3) ( 4 ), [(η3‐C3H5)(OC)3Re]2‐ (μ‐bis‐(diphenylphosphino)ferrocene) ( 8 ), (η1‐C3H5)Re(CO)3‐ (bpy) ( 14 ), of 19 , 20 and of (OC)3Re‐[Ph2P(CH2)3PPh2]Cl ( 16 ) were determined by X‐ray diffraction.  相似文献   

19.
The first 4π‐electron resonance‐stabilized 1,3‐digerma‐2,4‐diphosphacyclobutadiene [LH2Ge2P2] 4 (LH=CH[CHNDipp]2 Dipp=2,6‐iPr2C6H3) with four‐coordinate germanium supported by a β‐diketiminate ligand and two‐coordinate phosphorus atoms has been synthesized from the unprecedented phosphaketenyl‐functionalized N‐heterocyclic germylene [LHGe‐P=C=O] 2 a prepared by salt‐metathesis reaction of sodium phosphaethynolate (P≡C?ONa) with the corresponding chlorogermylene [LHGeCl] 1 a . Under UV/Vis light irradiation at ambient temperature, release of CO from the P=C=O group of 2 a leads to the elusive germanium–phosphorus triply bonded species [LHGe≡P] 3 a , which dimerizes spontaneously to yield black crystals of 4 as isolable product in 67 % yield. Notably, release of CO from the bulkier substituted [LtBuGe‐P=C=O] 2 b (LtBu=CH[C(tBu)N‐Dipp]2) furnishes, under concomitant extrusion of the diimine [Dipp‐NC(tBu)]2, the bis‐N,P‐heterocyclic germylene [DippNC(tBu)C(H)PGe]2 5 .  相似文献   

20.
A series of LZnX zinc/β‐ketoiminato complexes [L = CH3C(OH)?C(CH2CH?CH2)C(CH3)?NAr ( L1 ), CH3C(OH)?C(CH2CH2CN)C(CH3)?NAr ( L2 ), CH3C(OH)?C(CH2C6H5)C(CH3)?NAr ( L3 ), or CH3C(OH)?CHC(CH3)?NAr ( L4 ); Ar = 2,6‐iPr2C6H3; and initiation group X = alcoholate or acetate (for L1 ) or alcoholate (for L2 – L4 )] were synthesized, and their activities toward the copolymerization of carbon dioxide with cyclohexene oxide were determined. The 3‐position substituents on the β‐ketoiminato ligand backbone of the zinc/β‐ketoiminato complexes played an important role not only in the catalytic activity but also in the intrinsic viscosity, chemical composition, and refined microstructure of the resultant copolymers. The order of the catalytic activity of L1 ZnX with different initiation groups (X = OMe, OiPr, or OAc) was L1 Zn (OiPr) > L1 Zn (OMe) > L1 Zn (OAc), being the opposite of the order of the leaving ability of the initiation groups. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 6243–6251, 2006  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