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.
In toluene at reflux temperatures [Ru3(CO)12] and 7-SMe2-nido-7-CB10H12 give the charge-compensated cluster complex [1-SMe2-2,2-(CO)2-7,11-(μ-H)2-2,7,11-Ru2(CO)6-closo-2,1-RuCB10H8] (1) . Treatment of 1 with dppm in THF affords [1-SMe2-2,2-(CO)2-7,11-(μ-H)2-2,7,11-Ru2(μ -dppm)(CO)4-closo-2,1-RuCB10H8] (2) [dppm = bis(diphenylphosphino)methane; THF = tetrahydrofuran]. The latter complex on heating in THF with [ ]F yields the salt [ ][1-SMe-2,2-(CO)2-7,11-(μ-H)2-2,7,11-Ru2(μ -dppm)(CO)4-closo-2,1-RuCB10H8] (3). Reaction of 3 with [AuCl(PPh3)] and Tl[PF6] gives the neutral zwitterionic complex [1-S(Me)Au (PPh3)-2,2-(CO)2-7,11-(μ-H)2-2,7,11-Ru2(μ-dppm)(CO)4-closo-2,1-RuCB10H8] (4). The structures of 1, 3 and 4 were determined by single-crystal X-ray diffraction studies.*Dedicated to Professor F. Albert Cotton on the occasion of his 75th birthday, in appreciation of our long friendship and in recognition of his outstanding contributions to the study of complexes with metal–metal bonds.  相似文献   

2.
The ruthenium-tin complex, [Ru2(CO)4(SnPh3)2(μ-pyS)2] (1), the main product of the oxidative-addition of pySSnPh3 to Ru3(CO)12 in refluxing benzene, is [Ru(CO)2(pyS)(SnPh3)] synthon. It reacts with PPh3 to give [Ru(CO)2(SnPh3)(PPh3)(κ2-pyS)] (2) and further with Ru3(CO)12 or [Os3(CO)10(NCMe)2] to afford the butterfly clusters [MRu3(CO)12(SnPh3)(μ3-pyS)] (3, M=Ru; 4, M=Os). Direct addition of pySSnPh3 to [Os3(CO)10(NCMe)2] at 70 °C gives [Os3(CO)9(SnPh3)(μ3-pyS)] (5) as the only bimetallic compound, while with unsaturated [Os3(CO)83-PPh2CH2P(Ph)C6H4}(μ-H)] the previously reported [Os3(CO)8(μ-pyS)(μ-H)(μ-dppm)] (6) and the new bimetallic cluster [Os3(CO)7(SnPh3){μ-Ph2PCH2P(Ph)C6H4}(μ-pyS)[(μ-H)] (7) are formed at 110 °C. Compounds 1, 2, 4, 5 and 7 have been characterized by X-ray diffraction studies.  相似文献   

3.
Heterometallic triangular platinum–cobalt, palladium–cobalt and palladium–molybdenum clusters stabilized by one or two bridging diphosphine ligands such as Ph2PNHPPh2 (dppa) or (Ph2P)2NMe (dppaMe) or by mixed ligand sets Ph2PCH2PPh2 (dppm)/dppa have been prepared with the objectives of comparing the stability and properties of the clusters as a function of the short-bite diphosphine ligand used and of the metal carbonyl fragment they contain. Ligand redistribution reactions were observed during the purification of [Co2Pd(μ3-CO)(CO)4(μ-dppa)(μ-dppm)] (4) by column chromatography with the formation of [Co2Pd(μ3-CO)(CO)4(μ-dppm)2] and the dinuclear complex [(OC)2 Cl] (5). The latter was independently prepared by reaction of [Pd(dppa-P,P′)2](BF4)2 with Na[Co(CO)4]. Attempts to directly incorporate the ligand (Ph2P)2N(CH2)3Si(OMe)3 (dppaSi) into a cluster or to generate it by N-functionalization of coordinated dppa were unsuccessful, in contrast to results obtained recently with related clusters. The crystal structure of [Co2Pt(μ3-CO)(CO)6(μ-dppa)] (1) has been determined by X-ray diffraction.  相似文献   

4.
Treatment of closo-[Ru44-PPh)22-CO)(CO)10] with acetylene under ambient conditions leads to the insertion of the acetylene into the skeletal framework of the cluster and the formation of [Ru44-PPh){μ43-P(Ph)CHCH}(μ2-CO)(CO)10], the structure of which has been determined X-ray crystallographically.  相似文献   

5.
Mixed-chelate complexes of ruthenium have been synthesized using tridentate Schiff-base ligands (TDLs) derived from condensation of 2-aminophenol or 2-aminobenzoic acid with aldehydes (salicyldehyde, 2-pyridinecarboxaldehyde), and tmeda (tetramethylethylenediamine). [RuIII(hpsd)(tmeda)(H2O)]+ (1), [RuIII(hppc)(tmeda)(H2O)]2+ (2), [RuIII(cpsd)(tmeda)(H2O)]+ (3) and [RuIII(cppc)(tmeda)(H2O)]2+ (4) complexes (where hpsd2− = N-(hydroxyphenyl)salicylaldiminato); hppc = N-(2-hydroxyphenylpyridine-2-carboxaldiminato); cpsd2− = (N-(2-carboxyphenyl)salicylaldiminato); cppc = N-2-carboxyphenylpyridine-2-carboxaldiminato) were characterized by microanalysis, spectral (IR and UV–vis), conductance, magnetic moment and electrochemical studies. Complexes 14 catalyzed the epoxidation of cyclohexene, styrene, 4-chlorostyrene, 4-methylstyrene, 4-methoxystyrene, 4-nitrostyrene, cis- and trans-stilbenes effectively at ambient temperature using tert-butylhydroperoxide (t-BuOOH) as terminal oxidant. On the basis of Hammett correlation (log krel vs. σ+) and product analysis, a mechanism involving intermediacy of a [Ru–O–OBut] radicaloid species is proposed for the catalytic epoxidation process.  相似文献   

6.
Reaction of cis-[Mo(NCMe)2(CO)2(η5-L)][BF4] (L=C5H5 or C5Me5) with 1-acetoxybuta-1,3-diene gives the cationic complexes [Mo{η4-syn-s-cis-CH2CHCHCH(OAc)}(CO)2(η5-L)][BF4], which, on reaction with aqueous NaHCO3/CH2Cl2, afford good yields of the anti-aldehyde substituted complexes [Mo{η3-exo-anti-CH2CHCH(CHO)}(CO)2(η5-L)] 2 (L=C5Me5), 4 (L=C5H5)]. The corresponding η5-indenyl substituted complex 5 was prepared by protonation (HBF4·OEt2) of [Mo(η3-C3H5)(CO)2(η5-C9H7)] followed by addition of CH2=CHCH=CH(OAc) and hydrolysis (aq. NaHCO3/CH2Cl2). An X-ray crystallographic study of complex 2 confirmed the structure and showed that there is a contribution from a zwitterionic form involving donation of electron density from the molybdenum to the aldehyde carbonyl group. Treatment of 2 and 4, in methanol solution, with NaBH4 afforded the alcohols [Mo{η3-exo-anti-CH2CHCHCH2(OH)}(CO)2(η5-L)] [6 (L=C5H5), 8 (L=C5Me5)]; however, prolonged (30 h) reaction with NaBH4/MeOH surprisingly gave good yields of the methoxy-substituted complexes [Mo{η3-exo-anti-CH2CHCHCH2(OMe)}(CO)2(η5-L)] [7 (L=C5H5), 9 (L=C5Me5)], the structure of 7 being confirmed by single crystal X-ray crystallography. This methoxylation reaction can be explained by coordination of the hydroxyl group present in 6 and 8 onto B2H6 to form the potential leaving group HOBH3, which on ionisation affords [Mo(η4-exo-buta-1-3-diene)(CO)2(η5-L)]+ which is captured by reaction with OMe. Complex 8 is also formed in good yield on reaction of 2 with HBF4·OEt2 followed by treatment of the resulting cation [Mo{η4-exo-s-cis-syn-CH2CHCHCH(OH)}(CO)2(η5-C5Me5)][BF4] with Na[BH3CN]. Reaction of 4 with the Grignard reagents MeMgI, EtMgBr or PhMgCl afforded moderate yields of the alcohols [Mo{η3-exo-anti-CH2CHCHCH(OH)R}(CO)2(η5-C5H5)] [11 (R=Me), 12 (R=Et), 13 (R=Ph)]. Similarly, treatment of 2 with MeLi gave the corresponding alcohol 14. An attempt to carry out the Oppenauer oxidation [Al(OPr′)3/Me2CO] of 11 resulted in an elimination reaction and the formation of the η3-s-pentadienyl complex [Mo{η3-exo-anti-CH2CHCH(CHCH2)}(CO)2(η5-C5H5)], which was structurally identified by X-ray crystallography. Interestingly, oxidation of 6 with [Bu4nN][RuO4]/morpholine-N-oxide affords the aldehyde complex, 4 in good yield. Finally, reaction of 11 with [NO][BF4] followed by addition of Na2CO3 affords the fur-3-ene complex [Mo{η2-
(H)Me}(CO)(NO)(η5-C5H5)].  相似文献   

7.
Nickel(II), palladium(II), and platinum(II) complexes of 2-(3-mesitylimidazolylidenyl)pyrimidine (L), [Ni2(μ-Cl)2(L)4][Ag2Cl4] (3), [Ni2(μ-I)2(L)4][NiI(L)2(CH3CN)]2[Ag4I8] (4), [PdCl2(L)] (5), [PdI2(L)] (6), and [PtCl(L)2][AgCl2] (7) have been obtained from the carbene transfer reactions of [Ag(L)Cl] (2). These complexes have been fully characterized by spectroscopic methods and single-crystal X-ray structure analyses. The mono(carbene)palladium and bis(carbene)platinum complexes display normal square–planar structures. Nickel complexes 3 and 4 are rare examples of paramagnetic nickel(II) complexes of N-heterocyclic carbenes having octahedral geometry.  相似文献   

8.
The radical initiated reactions of Ru3(CO)12 with pyrazolyl substituted diphosphazanes Ph2PN(R)PPh(N2C3HMe2-3,5) [R = (S)-*CHMePh (1) or CHMe2 (2)] proceed via P–N(pyrazole) bond rupture resulting in the formation of phosphido clusters, [Ru3(CO)5sb-CO)23-N,N′-η111-N2C3HMe2-3,5){μ-P,P′-Ph2PN(R)PPh}] [R = (S)-*CHMePh (3) or CHMe2 (4)]. The pyrazolate moiety adopts an unusual triply bridging μ3111-mode of coordination in these clusters.  相似文献   

9.
《印度化学会志》2021,98(2):100023
The syntheses, structures and thermal reactions of [Ru3(CO)9{P(C4H3E)3}(μ-dppe)] (2, E = S; 3, E = O; dppe = 1,2-bis(diphenylphosphino)ethane) are described. These triphosphine-substituted clusters can be easily obtained in high yield from the Me3NO initiated room temperature reaction between [Ru3(CO)10(μ-dppe)] (1) and P(C4H3E)3. Both clusters have been structurally characterized which reveals that the functionalized phosphine P(C4H3E)3 is coordinated to the remote ruthenium atom using the phosphorus atom, while the NMR spectroscopic data indicate that both clusters are fluxional in solution mainly due to the ring-flipping process involving the dppe ligand which has been probed by VT NMR spectroscopy. Thermolysis of 2 at 66 °C affords 1 via P(C4H3S)3 dissociation, whilst that of 3 under similar experimental conditions also furnishes the diruthenium σ,π-furyl complex [Ru2(CO)6(μ,η2-C4H3O){μ-P(C4H3O)2] (4) in addition to 1.  相似文献   

10.
Half-titanocene is well-known as an excellent catalyst for the preparation of SPS (syndiotactic polystyrene) when activated with methylaluminoxane (MAO). Dinuclear half-sandwich complexes of titanium bearing a xylene bridge, (TiCl2L)2{(μ-η5, η5-C5H4-ortho-(CH2–C6H4–CH2)C5H4}, (4 (L = Cl), 7 (L = O-2,6-iPr2C6H3)) and (TiCl2L)2{(μ-η5, η5-C5H4-meta-(CH2–C6H4–CH2)C5H4} (5 (L = Cl), 8(L = O-2,6-iPr2C6H3)), have been successfully synthesized and introduced for styrene polymerization. The catalysts were characterized by 1H- and 13C NMR, and elemental analysis. These catalysts were found to be effective in forming SPS in combination with MAO. The activities of the catalysts with rigid ortho- and meta-xylene bridges were higher than those of catalysts with flexible pentamethylene bridges. The catalytic activity of four dinuclear half-titanocenes increased in the order of 4 < 5 < 7 < 8. This result displays that the meta-xylene bridged catalyst is more active than the ortho-xylene bridged and that the aryloxo group at the titanium center is more effective at promoting catalyst activity compared to the chloride group at the titanium center. Temperature and ratio of [Al]:[Ti] had significant effects on catalytic activity. Polymerizations were conducted at three different temperatures (25, 40, and 70 °C) with variation in the [Al]:[Ti] ratio from 2000 to 4000. It was observed that activity of the catalysts increased with increasing temperature, as well as higher [Al]:[Ti]. Different xylene linkage patterns (ortho and meta) were recognized to be a principal factor leading to the characteristics of the dinuclear catalyst due to its different spatial arrangement, causing dissimilar intramolecular interactions between the two active sites.  相似文献   

11.
The dinuclear gem-dithiolato bridged compounds [Rh2(μ-S2Cptn)(cod)2] (1) (CptnS22− = 1,1-cyclopentanedithiolato), [Rh2(μ-S2Chxn)(cod)2] (2) (ChxnS22− = 1,1-cyclohexanedithiolato), [Rh2(μ-S2CBn2)(cod)2] (3) (Bn2CS22− = 1,3-diphenyl-2,2-dithiolatopropane) and [Rh2(μ-S2CiPr2)(cod)2] (4) (iPr2CS22− = 2,4-dimethyl-2,2-dithiolatopentane) dissolved in toluene in the presence of monodentate phosphine or phosphite P-donor ligands under carbon monoxide/hydrogen (1:1) atmosphere are efficient catalysts for the hydroformylation of oct-1-ene under mild conditions (6.8 atm of CO/H2 and 80 °C). The influence of the gem-dithiolato ligand, the P-donor co-catalyst and the P/Rh ratio on the catalytic activity and selectivity has been explored. Aldehyde selectivities higher than 95% and turnover frequencies up to 245 h−1 have been obtained using P(OMe)3 as modifying ligand. Similar activity figures have been obtained using P(OPh)3 although the selectivities are lower. Regioselectivities toward linear aldehyde are in the range 75–85%. The performance of the catalytic systems [Rh2(μ-S2CR2)(CO)2(PPh3)2]/PPh3 has been found to be comparable to the systems [Rh2(μ-S2CR2)(cod)2] at the same P/Rh ratio. The system [Rh2(μ-S2CBn2)(cod)2] (3)/P(OPh)3 has been tested in the hydroformylation-isomerization of trans-oct-2-ene. Under optimized conditions up to 54% nonanal was obtained. Spectroscopic studies under pressure (HPNMR and HPIR) evidenced the formation of hydrido mononuclear species under catalytic conditions that are most probably responsible for the observed catalytic activity.  相似文献   

12.
Yang Fan  Phillip E. Fanwick  Tong Ren   《Polyhedron》2009,28(16):3654-3658
4-Vinylbenzoic acid reacted with Ru2(D(3,5-Cl2Ph)F)3(OAc)Cl and cis-Ru2(D(3,5-Cl2Ph)F)2(OAc)2Cl (D(3,5-Cl2Ph)F is N,N-bis(3,5-dichlorophenyl)formamidinate) to yield Ru2(D(3,5-Cl2Ph)F)3(4-vinylbenzoate)Cl (1) and cis-Ru2(D(3,5-Cl2Ph)F)2(4-vinylbenzoate)2Cl (2), respectively. Ru2(D(3,5-Cl2Ph)F)3(OAc)Cl reacted with 5-hexenoic acid and 6-heptenoic acid to afford Ru2(D(3,5-Cl2Ph)F)3(5-hexenoate)Cl (3) and Ru2(D(3,5-Cl2Ph)F)3(6-heptenoate)Cl (4), respectively. All new compounds were characterized using voltammetric and Vis–NIR spectroscopic techniques, and the structures of 1 and 2 were also established through X-ray single crystal diffractions.  相似文献   

13.
Treatment of [Ru3(CO)10(μ-dppm)] (4) [dppm = bis(diphenylphosphido)methane] with tetramethylthiourea at 66 °C gave the previously reported dihydrido triruthenium cluster [Ru3(μ-H)23-S)(CO)7(μ-dppm)] (5) and the new compounds [Ru33-S)2(CO)7(μ-dppm)] (6), [Ru33-S)(CO)73-CO)(μ-dppm)] (7) and [Ru33-S){η1-C(NMe2)2}(CO)63-CO)(μ-dppm)] (8) in 6%, 10%, 32% and 9% yields, respectively. Treatment of 4 with thiourea at the same temperature gave 5 and 7 in 30% and 10% yields, respectively. Compound 7 reacts further with tetramethylthiourea at 66 °C to yield 6 (30%) and a new compound [Ru33-S)21-C(NMe2)2}(CO)6(μ-dppm)] (9) (8%). Thermolysis of 8 in refluxing THF yields 7 in 55% yield. The reaction of 4 with selenium at 66 °C yields the new compounds [Ru33-Se)(CO)73-CO)(μ-dppm)] (10) and [Ru33-Se)(μ33-PhPCH2PPh(C6H4)}(CO)6(μ-CO)] (11) and the known compounds [Ru3(μ-H)23-Se)(CO)7(μ-dppm)] (12) and [Ru43-Se)4(CO)10(μ-dppm)] (13) in 29%, 5%, 2% and 5% yields, respectively. Treatment of 10 with tetramethylthiourea at 66 °C gives the mixed sulfur-selenium compounds [Ru33-S)(μ3-Se)(CO)7(μ-dppm)] (14) and [Ru33-S)(μ3-Se){η1-C(NMe2)2}(CO)6(μ-dppm)] (15) in 38% and 10% yields, respectively. The single-crystal XRD structures of 6, 7, 8, 10, 14 and 15 are reported.  相似文献   

14.
Thermal reaction of [Ru2(CO)6(μ-PFu2)(μ-η12-Fu)] (Fu=2-furyl) with (9-anthracenyl)diphenylphosphine (AnPPh2) produces a novel diruthenium complex [Ru2(CO)5(μ-PFu2)(μ-η112-C14H8PPh2)] (1) in good yield whereas the corresponding reaction between [(μ-H)4Ru4(CO)12] and AnPPh2 gives [HRu(CO)3(PPh2C14H8)][(μ-H)4Ru4(CO)11(AnPPh2)] (2). Both compounds 1 and 2 were fully characterized by spectroscopic methods and their X-ray crystal structures were determined. For 1, initial coordination of the PPh2 functionality at the Ru atom is accompanied by cyclometalation of the anthracenyl ring to form a Ru–C σ bond together with concomitant formation of a π bond to the adjacent Ru center and loss of the furyl ligand. The formation of 2 involves the cleavage of two Ru–Ru bonds, and the making of a Ru–P bond, followed by orthometalation of the anthracenyl ring. The optical absorption and emission spectra of 1 were recorded and the results were correlated to the DFT calculations.Dedicated to Professor F. Albert Cotton on the occasion of his 75th birthday.  相似文献   

15.
2-Methylthiothiophene (2-MeSC4H3S) oxidatively adds to [Os3(CO)10(MeCN)2] with cleavage of the C–H bond at the 3-position to give [Os3(μ-H)(μ-MeSC4H2S)(CO)10] 1, the X-ray structure of which shows that the MeS group is coordinated to osmium through the S atom while the thiophene ring is coordinated to osmium through the 3-carbon atom. Only one invertomer at sulfur is observed in solution and in the crystal the Me group is exo. Thermal treatment of 1 in the dark gives only one product, [Os3(μ-H)(μ3-MeSC4H2S)(CO)9] 2 (X-ray structure), derived by loss of a CO from the Os(CO)4 unit of 1 with concomitant η2-coordination of the thiophene ring of bridging MeSC4H2S at the third metal atom. Whereas thermal reaction in the dark leads only to C–H cleavage products, visible irradiation at room temperature leads to various products derived by migration of the MeS group. Thus thermal treatment of 1 in daylight for 2 h gave 2, together with an isomer 3. Cluster 2 converts at room temperature to 3 in daylight while thermal treatment of 2 in the dark (125°C) gave no reaction. Isomer 3 of [Os3(μ-H)(μ3-MeSC4H2S)(CO)9] (X-ray structure) is closely related to 2 except that the MeS group and the Os–C σ-bond have interchanged sites at the thiophene ring between the 2- and the 3-positions. Visible irradiation of 1 at room temperature for 3 days in daylight gave further chemical change leading to two bridging thienyl clusters, [Os3(μ-MeS)(μ-2-C4H2S)(CO)10] 4 and [Os3(μ-MeS)(μ3-2-C4H2S)(CO)9] 5. Cluster 5 is the ultimate product of daylight irradiation of any of the clusters 1 to 4.  相似文献   

16.
《Polyhedron》2001,20(15-16):2011-2018
The reaction behavior of the 48e-clusters [Ru3(CO)8(μ-H)2(μ-PR2)2] (R=But, 1a; R=Cy, 1b) towards phosphine ligands has been studied. Whereas 1a reacts spontaneously with many phosphines at room temperature, a lack of reactivity for 1b under similar conditions is observed. Thus 1a reacts with dppm (Ph2PCH2PPh2) to the known 46e-cluster [Ru3(μ-CO)(CO)43-H)(μ-H)(μ-PBut2)2(μ-dppm)] (2a), and the reaction of 1a with dppe (Ph2PC2H4PPh2) yields analogously [Ru3(μ-CO)(CO)43-H)(μ-H)(μ-PBut2)2(μ-dppe)] (3). Reactions of 1a with dmpm (Me2PCH2PMe2), dmpe (Me2PC2H4PMe2) and PBun3, respectively, gave in each case a mixture of products which could not be characterized. Contrary to the reaction behavior at room temperature, 1b reacts with phosphines in THF under reflux yielding the novel complexes [Ru3(CO)6(μ-H)2(μ-PCy2)2L2] (L=Cy2PH, 4a; L=But2PH, 4b; L=Ph2PH, 4c; L=P(OEt)3, 4d). 4a is also obtained directly by the reaction of [Ru3(CO)12] with an excess of Cy2PH. The molecular structure of 4a has been determined by a single-crystal X-ray analysis. Moreover, the thermolysis of 1a in octane affords [Ru3(CO)8(μ-H)23-PBut)(But2PH)] (6) as the main product, and the thermolysis of [Ru3(CO)9(But2PH)(μ-dppm)] (7) yields 2a to a considerable extent. Treatment of 1a with carbon tetrachloride leads to [Ru3(CO)7(μ-H)(μ-PBut2)2(μ-Cl)] (8) as the main product.  相似文献   

17.
Thermal reaction of [Ru3(CO)12] with PH2Mes (Mes = mesityl) in refluxing toluene afforded mesitylphosphinidene-capped ruthenium carbonyl clusters, [Ru3(CO)9(μ-H)23-PMes)] (1), [Ru3(CO)8(PH2Mes)(μ-H)23-PMes)] (2), [Ru3(CO)93-PMes)2] (3), [Ru4(CO)10(μ-CO)(μ4-PMes)2] (4), and [Ru5(CO)10H24-PMes)(μ3-PMes)2] (5). All products were fully characterized and structurally confirmed by X-ray crystal structure analysis. Complexes 2-4 were also obtained in high yields by stepwise reaction starting from 1. Fluxional behavior of carbonyl groups was observed in case of 4. Complex 5 reveals a new type of skeletal structure, bicapped-octahedron having μ3- and μ4-phosphinidene ligands at the capping positions. Similar reaction of [Os3(CO)12] with PH2Mes yielded a phosphido-bridged osmium cluster [Os3(CO)10(μ-H)(μ-PHMes)] (6) and a phosphinidene-capped cluster [Os3(CO)9(μ-H)23-PMes)] (7).  相似文献   

18.
Reaction of [Ru3(CO)10(μ-dppm)] (1) with H2S at 66 °C affords high yields of the sulfur-capped dihydride [Ru3(CO)7(μ-H)2(μ-dppm)(μ3-S)] (2), formed by oxidative-addition of both hydrogen-sulfur bonds. Hydrogenation of [Ru3(CO)7(μ-dppm)(μ3-CO)(μ3-S)] (3) at 110 °C also gives 2 in similar yields, while hydrogenation of [Ru3(CO)7(μ-dppm)(μ3-CO)(μ3-Se)] (4) affords [Ru3(CO)7(μ-H)2(μ-dppm)(μ3-Se)] (5) in 85% yield. The molecular structures of 2 and 5 reveal that the diphosphine and one hydride simultaneously bridge the same ruthenium-ruthenium edge with the second hydride spanning one of the non-bridged edges. Both 2 and 5 are fluxional at room temperature being attributed to hydride migration between the non-bridged edges. Addition of HBF4 to 2 affords the cationic trihydride [Ru3(CO)7(μ-H)3(μ-dppm)(μ3-S)][BF4] (6) in which the hydrides are non-fluxional due to the blocking of the free ruthenium-ruthenium edge.  相似文献   

19.
The article describes the synthesis and single-crystal X-ray analysis of two sulfato and one thiocyanato copper(II) complex with 2-acetylpyridine S-methylisothiosemicarbazone (HL) of the formulae [Cu(HL)SO4(H2O)]·H2O (1), [Cu2(HL)2(μ-SO4)2]·2H2O (2) and [Cu(HL)(NCS)(SCN)] (3), as well as the structure of the protonated ligand H2L+I. Complexes 1 and 2 were obtained from the reaction of aqueous/methanolic CuSO4·5H2O and ethanolic/methanolic H2L+I solutions, respectively. Complex 3 was synthesized by the reaction of methanolic solutions of Cu(ClO4)2·6H2O, the ligand and NH4SCN, with the addition of triethyl orthoformate. All three complexes have a slightly deformed square-pyramidal structure (τav = 0.15) with the tridentate NNN neutral ligand in the basal plane. In complexes 1 and 3 the apical position is occupied by the oxygen atom of the monodentate SO4 group, or the sulfur atom of the SCN group. Thanks to the hydrogen bonds, complex 3 may be thought of as having a pseudo-dimeric structure. In the authentic centrosymmetric dimer 2, the oxygen atoms of both SO4 groups occupy also the apical position of both coordination polyhedra, as well as an equatorial position. Complexes 1 and 3 have μeff values characteristic of magnetically isolated mononuclear Cu(II) complexes. In contrast to them, complex 2 has a μeff value of 1.57 BM, which is in agreement with its dinuclear structure. All the complexes, in addition to the X-ray analysis and magnetic measurements, were characterized by IR and UV–Vis spectroscopy and by thermal analysis.  相似文献   

20.
Multinuclear NMR data (13C, 31P, 13C–{31P}, 13C–{103Rh} and 31P–{103Rh}) for a series of mono- and di-substituted derivatives of Rh6(CO)16 containing neutral two electron donor ligands [Rh6(CO)15L, (L=NCMe, py, cyclooctene, PPh3, P(OPh)3,1/2(μ2,η1:η1-dppe)); Rh6(CO)14(LL), (LL=cis-CH2=CMe-CMe=CH2, dppm, dppe, (P(OPh)3)2)] are reported; these data show that the solid state structure is maintained in solution. Detailed assignments of the 13CO NMR spectra of Rh6(CO)15(PPh3) and Rh6(CO)14(dppm) clusters have been made on the basis 13C–{103Rh} double resonance measurements and the specific stereochemical features of the observed long range couplings in these clusters have been studied. The stereochemical dependence of 3J(P–C) for terminal carbonyl ligands is discussed and the values of 3J(P–C) are found to be mainly dependent on the bond angles in the P–Rh–Rh–C fragment; these data enable the fine structure of the complex multiplets in the 13C–{1H} and 31P–{1H} NMR spectra of Rh6(CO)14 (dppm) to be simulated. Variable temperature 13C–{1H} NMR measurements on Rh6(CO)15(PPh3) reveal the carbonyl ligands in this complex to be fluxional. The fluxional process involves exchange of all the CO ligands except the two terminal CO's associated with the rhodium trans to the substituted rhodium and can be explained by a simple oscillation of the PPh3 on the substituted rhodium atom aided by concomitant exchange of the unique terminal CO on this rhodium with adjacent μ3-CO's.  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