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.
A series of ruthenium alkenylacetylide complexes trans-[Ru{C≡CC(=CH2)R}Cl(dppe)2] (R=Ph ( 1 a ), cC4H3S ( 1 b ), 4-MeS-C6H4 ( 1 c ), 3,3-dimethyl-2,3-dihydrobenzo[b]thiophene (DMBT) ( 1 d )) or trans-[Ru{C≡C-cC6H9}Cl(dppe)2] ( 1 e ) were allowed to react with the corresponding propargylic alcohol HC≡CC(Me)R(OH) (R=Ph ( A ), cC4H3S ( B ), 4-MeS-C6H4 ( C ), DMBT ( D ) or HC≡C-cC6H10(OH) ( E ) in the presence of TlBF4 and DBU to presumably give alkenylacetylide/allenylidene intermediates trans-[Ru{C≡CC(=CH2)R}{C=C=C(Me)}(dppe)2]PF6 ([ 2 ]PF6). These complexes were not isolated but deprotonated to give the isolable bis(alkenylacetylide) complexes trans-[Ru{C≡CC(=CH2)R}2(dppe)2] (R=Ph ( 3 a ), cC4H3S ( 3 b ), 4-MeS-C6H4 ( 3 c ), DMBT ( 3 d )) and trans-[Ru{C≡C-cC6H9}2(dppe)2] ( 3 e ). Analogous reactions of trans-[Ru(CH3)2(dmpe)2], featuring the more electron-donating 1,2-bis(dimethylphosphino)ethane (dmpe) ancillary ligands, with the propargylic alcohols A or C and NH4PF6 in methanol allowed isolation of the intermediate mixed alkenylacetylide/allenylidene complexes trans-[Ru{C≡CC(=CH2)R}{C=C=C(Me)}(dmpe)2]PF6 (R=Ph ([ 4 a ]PF6), 4-MeS-C6H4 ([ 4 c ]PF6). Deprotonation of [ 4 a ]PF6 or [ 4 c ]PF6 gave the symmetric bis(alkenylacetylide) complexes trans-[Ru{C≡CC(=CH2)R}2(dmpe)2] (R=Ph ( 5 a ), 4-MeS-C6H4 ( 5 c )), the first of their kind containing the dmpe ancillary ligand sphere. Attempts to isolate bis(allenylidene) complexes [Ru{C=C=C(Me)R}2(PP)2]2+ (PP=dppe, dmpe) from treatment of the bis(alkenylacetylide) species 3 or 5 with HBF4 ⋅ Et2O were ultimately unsuccessful.  相似文献   

2.
Areneruthenium(II) compounds [Ru(p‐cym)Cl2{κPiPrP(CH2CH2OMe)2}], 3 , and [Ru(arene)Cl2{κP‐RP(CH2CO2Me)2}] 4 – 7 (arene=p‐cym (=1‐methyl‐4‐isopropylbenzene), mes (=1,3,5‐trimethylbenzene); R=iPr, tBu) were prepared from the dimers [Ru(arene)Cl2]2 and the corresponding functionalized phosphine. Treatment of 6 and 7 with 1 equiv. of AgPF6 affords the monocationic complexes [Ru(mes)Cl{κ2P,O‐RP(CH2C(O)OMe)(CH2CO2Me)}]PF6, 10 and 11 , while the related reaction of 5 – 7 with 2 equiv. of AgPF6 produces the dicationic compounds [Ru(p‐cym){κ3P,O,O‐tBuP(CH2C(O)OMe)2}](PF6)2 ( 12 ) and [Ru(mes){κ3P,O,O‐RP(CH2C(O)OMe)2}](PF6)2, 13 and 14 . Partial hydrolysis of one hexafluorophosphate anion of 12 – 14 leads to the formation of [Ru(arene){κ2P,O‐RP(CH2C(O)OMe)(CH2CO2Me)}(κO‐O2PF2)]PF6, 15 – 17 , of which 17 (arene=mes; R=tBu) has been characterized by X‐ray crystallography. Compounds 13 and 14 react with 2 equiv. of KOtBu in tBuOH/toluene to give the unsymmetrical complexes [Ru(mes){κ3P,C,O‐RP(CHCO2Me)(CH=C(O)OMe)}], 18 and 19 , containing both a five‐membered phosphinoenolate and a three‐membered phosphinomethanide ring. The molecular structure of compound 18 has been determined by X‐ray structure analysis. The neutral bis(carboxylate)phosphanidoruthenium(II) complexes [Ru(arene){κ3P,O,O‐RP(CH2C(O)O)2}], 20 – 23 are obtained either by hydrolysis of 18 and 19 , or by stepwise treatment of 4 and 5 with KOtBu and basic Al2O3. Novel tripodal chelating systems are generated via insertion reactions of 19 with PhNCO and PhNCS.  相似文献   

3.
[(η-C5H5)Ru{Ph2PCHRCHR′PPh2}({C(OCH3)CH2C6H5})]PF6 (where R, R′ = H or CH3) reacts with LiAlH4 in THF at ?80° C to give the corresponding 2-phenylethyl complexes, which have an antiperiplanar conformation around the H2CCH2 bond in solution; the reaction takes place with retention of configuration at the ruthenium atom.  相似文献   

4.
The reaction of [(η5‐L3)Ru(PPh3)2Cl], where; L3 = C9H7 ( 1 ), C5Me5 (Cp*) ( 2 ) with acetonitrile in the presence of [NH4][PF6] yielded cationic complexes [(η5‐L3)Ru(PPh3)2(CH3CN)][PF6]; L3= C9H7 ([3]PF6) and L3 = C5Me5 ([4]PF6), respectively. Complexes [3]PF6 and [4]PF6 reacts with some polypyridyl ligands viz, 2,3‐bis (α‐pyridyl) pyrazine (bpp), 2,3‐bis (α‐pyridyl) quinoxaline (bpq) yielding the complexes of the formulation [(η5‐L3)Ru(PPh3)(L2)]PF6 where; L3 = C9H7, L2 = bpp, ([5]PF6), L3 = C9H7, L2 = bpq, ([6]PF6); L3 = C5Me5, L2 = bpp, ([7]PF6) and bpq, ([8]PF6), respectively. However reaction of [(η5‐C9H7)Ru(PPh3)2(CH3CN)][PF6] ([3]PF6) with the sterically demanding polypyridyl ligands, viz. 2,4,6‐tris(2‐pyridyl)‐1,3,5‐triazine (tptz) or tetra‐2‐pyridyl‐1,4‐pyrazine (tppz) leads to the formation of unexpected complexes [Ru(PPh3)2(L2)(CH3CN)][PF6]2; L2 = tppz ([9](PF6)2), tptz ([11](PF6)2) and [Ru(PPh3)2(L2)Cl][PF6]; L2 = tppz ([10]PF6), tptz ([12]PF6). The complexes were isolated as their hexafluorophosphate salts. They have been characterized on the basis of micro analytical and spectroscopic data. The crystal structures of the representative complexes were established by X‐ray crystallography.  相似文献   

5.
Two ruthenium acetylide complexes [Ru]?C≡C?C≡C?C(OR)(C3H5)2 ( 2 , R=H and 2 a , R=CH3; [Ru]=Cp(PPh3)2Ru) each with two cyclopropyl rings were synthesized from TMS?C≡C?C≡C?C(OH)(C3H5)2 ( 1 ; TMS=trimethylsilyl). Treatments of 2 and 2 a with allyl halide in the presence of KPF6 afforded the vinylidene complexes 3 and 3 a , respectively. When NH4PF6 was used, instead of KPF6, additional ring‐opening reaction took place on one of the three‐membered ring. Treatment of [Ru]Cl with 1,3‐butadiyne ( 6 ), bearing an epoxide ring, afforded acetylide complex 7 with a furyl ring. Treatment of 2 a with Ph3CPF6 presumably afforded pentatetraenylidene complex {[Ru]=C=C=C=C=C(C3H5)2}[PF6] ( 10 ), which was not isolated. Additions of various alcohols in a solution of 10 generated a number of disubstituted allenylidene complexes {[Ru]=C=C=C(OR)?C=C(C3H5)2}[PF6] ( 11 ). Treatment of 11 with K2CO3 afforded the acetylide complex 12 bearing a carbonyl group, characterized by single X‐ray diffraction analysis. Addition of a primary amine to 10 caused cleavage of the farthermost C=C bond and several allenylidene complexes {[Ru]=C=C=C(Me)(NHR)}[PF6] ( 18 ) were isolated.  相似文献   

6.
The bonding modes of the ligand di‐2‐pyridyl ketoxime towards half‐sandwich arene ruthenium, Cp*Rh and Cp*Ir complexes were investigated. Di‐2‐pyridyl ketoxime {pyC(py)NOH} react with metal precursor [Cp*IrCl2]2 to give cationic oxime complexes of the general formula [Cp*Ir{pyC(py)NOH}Cl]PF6 ( 1a ) and [Cp*Ir{pyC(py)NOH}Cl]PF6 ( 1b ), for which two coordination isomers were observed by NMR spectroscopy. The molecular structures of the complexes revealed that in the major isomer the oxime nitrogen and one of the pyridine nitrogen atoms are coordinated to the central iridium atom forming a five membered metallocycle, whereas in the minor isomer both the pyridine nitrogen atoms are coordinated to the iridium atom forming a six membered metallacyclic ring. Di‐2‐pyridyl ketoxime react with [(arene)MCl2]2 to form complexes bearing formula [(p‐cymene)Ru{pyC(py)NOH}Cl]PF6 ( 2 ); [(benzene)Ru{pyC(py)NOH}Cl]PF6 ( 3 ), and [Cp*Rh{pyC(py)NOH}Cl]PF6 ( 4 ). In case of complex 3 the ligand coordinates to the metal by using oxime nitrogen and one of the pyridine nitrogen atoms, whereas in complex 4 both the pyridine nitrogen atoms are coordinated to the metal ion. The complexes were fully characterized by spectroscopic techniques.  相似文献   

7.
The reaction of [Ru(OH2)2(RaaiR′)2]2+ [RaaiR′ = 1-alkyl-2-(arylazo)imidazole, p-R–C6H4–N=N–C3H2–NN(1)–R′, R=H (1), Me (2), Cl (3); R′ = Me (a), Et (b), CH2Ph (c)] with 8-quinolinol (HQ) in acetone solution followed by the addition of NH4PF6 afforded violet, mixed ligand complexes of composition [Ru(Q)(RaaiR′)2](PF6). The structure of [Ru(Q)(MeaaiMe)2](PF6) (2a) has been confirmed by X-ray diffraction studies. Solution electronic spectra exhibit a strong MLCT band at 560–580?nm in MeCN. Cyclic voltammogrames show a Ru(III)/Ru(II) couple at 1.0–1.1?V versus SCE along with three successive ligand reductions. The electronic properties are correlated with EHMO results.  相似文献   

8.
Cationic substitutionally inert cyclometalated ruthenium (II) and osmium (II) complexes, ([Mt(o‐C6H4‐2‐py)(LL)2]PF6), where LL‐1,10‐phenanthroline (phen) or 2,2′‐bipyridine (bipy), were used for radical polymerization of styrene. Gradual modification of the complexes within the series allowed comparison of the catalytic activity and the redox properties. There was no correlation between the reducing powers of the complexes and their catalytic activities. The osmium compound of the lowest reduction potential was not active. All the ruthenium complexes catalyzed the polymerization of styrene in a controlled manner; but the level of control and the catalytic activity were different under the same polymerization conditions. [Ru(o‐C6H4‐2‐py)(phen)2]PF6 demonstrated the best catalytic performance though its redox potential was the highest. It catalyzed the “living” polymerization with a reasonable rate at a catalyst‐to‐initiator ratio of 0.1. 1 equiv. of Al(OiPr)3 accelerated the polymerization and improved the control, but higher amount of Al(OiPr)3 did not speed up the polymerization and moved the process into the uncontrollable regime. Under the most optimal conditions, the controlled polymerization occurs fast without any additive and the catalyst degradation. Added free ligands inhibited the polymerization suggesting that the catalytically active ruthenium intermediates are generated via the reversible dechelation of bidentate phen or bipy ligands. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 3814–3828, 2009  相似文献   

9.
A series of ruthenium hydride compounds containing substituted bidentate pyrrole‐imine ligands were synthesized and characterized. Reacting RuHCl(CO)(PPh3)3 with one equivalent of [C4H3NH(2‐CH=NR)] in ethanol in the presence of KOH gave compounds {RuH(CO)(PPh3)2[C4H3N(2‐CH=NR)]} where trans‐Py‐Ru‐H 1, R = CH2CH2C6H9; cis‐Py‐Ru‐H 2, R = Ph‐2‐Me; and cis‐Py‐Ru‐H 3, R = C6H11. Heating trans‐Py‐Ru‐H 1 in toluene at 70°C for 12 hr resulted a thermal conversion of the trans‐Py‐Ru‐H 1 into its cis form, {RuH(CO)(PPh3)2[C4H3N(2‐CH=NCH2CH2C6H9)]} (cis‐Py‐Ru‐H 1) in very high yield. The 1H NMR spectra of trans‐Py‐Ru‐H 1, cis‐Py‐Ru‐H 2, cis‐Py‐Ru‐H 3, and cis‐Py‐Ru‐H 1 all show a typical triplet at ca. δ–11 for the hydride. The trans and cis form indicate the relative positions of pyrrole ring and hydride. The geometries of trans‐Py‐Ru‐H 1, cis‐Py‐Ru‐H 1, and cis‐Py‐Ru‐H 3 are relatively similar showing typical octahedral geometries with two PPh3 fragments arranged in trans positions.  相似文献   

10.
Thermolysis of [Cp*Ru(PPh2(CH2)PPh2)BH2(L2)] 1 (Cp*=η5‐C5Me5; L=C7H4NS2), with terminal alkynes led to the formation of η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)B{R‐C=CH2}(L)2] ( 2 a – c ) and η2‐vinylborane complexes [Cp*Ru(R‐C=CH2)BH(L)2] ( 3 a – c ) ( 2 a , 3 a : R=Ph; 2 b , 3 b : R=COOCH3; 2 c , 3 c : R=p‐CH3‐C6H4; L=C7H4NS2) through hydroboration reaction. Ruthenium and the HBCC unit of the vinylborane moiety in 2 a – c are linked by a unique η4‐interaction. Conversions of 1 into 3 a – c proceed through the formation of intermediates 2 a – c . Furthermore, in an attempt to expand the library of these novel complexes, chemistry of σ‐borane complex [Cp*RuCO(μ‐H)BH2L] 4 (L=C7H4NS2) was investigated with both internal and terminal alkynes. Interestingly, under photolytic conditions, 4 reacts with methyl propiolate to generate the η4‐σ,π‐borataallyl complexes [Cp*Ru(μ‐H)BH{R‐C=CH2}(L)] 5 and [Cp*Ru(μ‐H)BH{HC=CH‐R}(L)] 6 (R=COOCH3; L=C7H4NS2) by Markovnikov and anti‐Markovnikov hydroboration. In an extension, photolysis of 4 in the presence of dimethyl acetylenedicarboxylate yielded η4‐σ,π‐borataallyl complex [Cp*Ru(μ‐H)BH{R‐C=CH‐R}(L)] 7 (R=COOCH3; L=C7H4NS2). An agostic interaction was also found to be present in 2 a – c and 5 – 7 , which is rare among the borataallyl complexes. All the new compounds have been characterized in solution by IR, 1H, 11B, 13C NMR spectroscopy, mass spectrometry and the structural types were unequivocally established by crystallographic analysis of 2 b , 3 a – c and 5 – 7 . DFT calculations were performed to evaluate possible bonding and electronic structures of the new compounds.  相似文献   

11.
Addition of [C7H7][PF6] to iron, ruthenium or osmium alkynyl complexes has given eight cationic cycloheptatrienylvinylidene derivatives [M{C C(C7H7)R}(L)2 (η-C5H5)][PF6] (M = Fe, Ru or Os; R = Me, Pr, Ph or C6F5; L = PPh3, L2 = dppm or dppe; but not all combinations). With Fe(C2Ph)(CO)2(η-C5H5), only [Fe(CO)2(thf)(η-C5H5)][PF6] was obtained. Reactions of the new complexes are characterised by loss of the C7H7 group. The NMR spectra and FAB mass spectra are described in detail.  相似文献   

12.
A new series of monoselenoquinone and diselenoquinone π complexes, [(η6p‐cymene)Ru(η4‐C6R4SeE)] (R=H, E=Se ( 6 ); R=CH3, E=Se ( 7 ); R=H, E=O ( 8 )), as well as selenolate π complexes [(η6p‐cymene)Ru(η5‐C6H3R2Se)][SbF6] (R=H ( 9 ); R=CH3 ( 10 )), stabilized by arene ruthenium moieties were prepared in good yields through nucleophilic substitution reactions from dichlorinated‐arene and hydroxymonochlorinated‐arene ruthenium complexes [(η6p‐cymene)Ru(C6R4XCl)][SbF6]2 (R=H, X=Cl ( 1 ); R=CH3, X=Cl ( 2 ); R=H, X=OH ( 3 )) as well as the monochlorinated π complexes [(η6p‐cymene)Ru(η5‐C6H3R2Cl)][SbF6]2 (R=H ( 4 ); R=CH3 ( 5 )). The X‐ray crystallographic structures of two of the compounds, [(η6p‐cymene)Ru(η4‐C6Me4Se2)] ( 7 ) and [(η6p‐cymene)Ru(η4‐C6H4SeO)] ( 8 ), were determined. The structures confirm the identity of the target compounds and ascertain the coordination mode of these unprecedented ruthenium π complexes of selenoquinones. Furthermore, these new compounds display relevant cytotoxic properties towards human ovarian cancer cells.  相似文献   

13.
A new series of mono- and bis-alkynyl CoIII(TIM) complexes (TIM=2,3,9,10-tetramethyl-1,4,8,11-tetraazacyclotetradeca-1,3,8,10-tetraene) is reported herein. The trans-[Co(TIM)(C2R)Cl]+ complexes were prepared from the reaction between trans-[Co(TIM)Cl2]PF6 and HC2R (R=tri(isopropyl)silyl or TIPS ( 1 ), -C6H4-4-tBu ( 2 ), -C6H4-4-NO2 ( 3 a ), and N-mesityl-1,8-naphthalimide or NAPMes ( 4 a )) in the presence of Et3N. The intermediate complexes of the type trans-[Co(TIM)(C2R)(NCMe)](PF6)(OTf), 3 b and 4 b , were obtained by treating 3 a and 4 a , respectively, with AgOTf in CH3CN. Furthermore, bis-alkynyl trans-[Co(TIM)(C2R)2]PF6 complexes, 3 c and 4 c , were generated following a second dehydrohalogenation reaction between 3 b and 4 b , respectively, and the appropriate HC2R in the presence of Et3N. These new complexes have been characterized using X-ray diffraction ( 2 , 3 a , 4 a , and 4 c ), IR, 1H NMR, UV/Vis spectroscopy, fluorescent spectroscopy ( 4 c ), and cyclic voltammetry.  相似文献   

14.
Reactions of [Ru{C=C(H)-1,4-C6H4C≡CH}(PPh3)2Cp]BF4 ([ 1 a ]BF4) with hydrohalic acids, HX, results in the formation of [Ru{C≡C-1,4-C6H4-C(X)=CH2}(PPh3)2Cp] [X=Cl ( 2 a-Cl ), Br ( 2 a-Br )], arising from facile Markovnikov addition of halide anions to the putative quinoidal cumulene cation [Ru(=C=C=C6H4=C=CH2)(PPh3)2Cp]+. Similarly, [M{C=C(H)-1,4-C6H4-C≡CH}(LL)Cp ]BF4 [M(LL)Cp’=Ru(PPh3)2Cp ([ 1 a ]BF4); Ru(dppe)Cp* ([ 1 b ]BF4); Fe(dppe)Cp ([ 1 c ]BF4); Fe(dppe)Cp* ([ 1 d ]BF4)] react with H+/H2O to give the acyl-functionalised phenylacetylide complexes [M{C≡C-1,4-C6H4-C(=O)CH3}(LL)Cp’] ( 3 a – d ) after workup. The Markovnikov addition of the nucleophile to the remote alkyne in the cations [ 1 a–d ]+ is difficult to rationalise from the vinylidene form of the precursor and is much more satisfactorily explained from initial isomerisation to the quinoidal cumulene complexes [M(=C=C=C6H4=C=CH2)(LL)Cp’]+ prior to attack at the more exposed, remote quaternary carbon. Thus, whilst representative acetylide complexes [Ru(C≡C-1,4-C6H4-C≡CH)(PPh3)2Cp] ( 4 a ) and [Ru(C≡C-1,4-C6H4-C≡CH)(dppe)Cp*] ( 4 b ) reacted with the relatively small electrophiles [CN]+ and [C7H7]+ at the β-carbon to give the expected vinylidene complexes, the bulky trityl ([CPh3]+) electrophile reacted with [M(C≡C-1,4-C6H4-C≡CH)(LL)Cp’] [M(LL)Cp’=Ru(PPh3)2Cp ( 4 a ); Ru(dppe)Cp* ( 4 b ); Fe(dppe)Cp ( 4 c ); Fe(dppe)Cp* ( 4 d )] at the more exposed remote end of the carbon-rich ligand to give the putative quinoidal cumulene complexes [M{C=C=C6H4=C=C(H)CPh3}(LL)Cp’]+, which were isolated as the water adducts [M{C≡C-1,4-C6H4-C(=O)CH2CPh3}(LL)Cp’] ( 6 a–d ). Evincing the scope of the formation of such extended cumulenes from ethynyl-substituted arylvinylene precursors, the rather reactive half-sandwich (5-ethynyl-2-thienyl)vinylidene complexes [M{C=C(H)-2,5-cC4H2S-C≡CH}(LL)Cp’]BF4 ([ 7 a – d ]BF4 add water readily to give [M{C≡C-2,5-cC4H2S-C(=O)CH3}(LL)Cp’] ( 8 a – d )].  相似文献   

15.
Piano‐stool‐shaped platinum group metal compounds, stable in the solid state and in solution, which are based on 2‐(5‐phenyl‐1H‐pyrazol‐3‐yl)pyridine ( L ) with the formulas [(η6‐arene)Ru( L )Cl]PF6 {arene = C6H6 ( 1 ), p‐cymene ( 2 ), and C6Me6, ( 3 )}, [(η6‐C5Me5)M( L )Cl]PF6 {M = Rh ( 4 ), Ir ( 5 )}, and [(η5‐C5H5)Ru(PPh3)( L )]PF6 ( 6 ), [(η5‐C5H5)Os(PPh3)( L )]PF6 ( 7 ), [(η5‐C5Me5)Ru(PPh3)( L )]PF6 ( 8 ), and [(η5‐C9H7)Ru(PPh3)( L )]PF6 ( 9 ) were prepared by a general method and characterized by NMR and IR spectroscopy and mass spectrometry. The molecular structures of compounds 4 and 5 were established by single‐crystal X‐ray diffraction. In each compound the metal is connected to N1 and N11 in a k2 manner.  相似文献   

16.
η6-o-Chlorotoluene-η5-cyclopentadienyliron hexafluorophosphate undergoes nucleophilic substitution of the chlorine atom with anions generated (K2CO3/DMF) from methyl thioglycolate, diethyl malonate, dimethyl malonate, methyl acetoacetate and 2,4-pentanedione. The compounds prepared were o-CH3C6H4SCH2CO2CH3FeCp+PF6, o-CH3C6H4CH(CO2C2H5)2FeCp+PF6, o-CH3C6H4CH(CO2CH3)2FeCp+PF6, o-CH3C6H4CH(COCH3)CO2CH3FeCp+PF6 and o-CH3C6H4CH2COCH3FeCp+PF6 . Similarly, the reaction of diethyl malonate, dimethyl malonate, methyl acetoacetate anions and methylamine with η6-2,6-dichlorotoluene-η5-cyclopentadienyliron hexafluorophosphate yielded monosubstitution of one of the chloro groups. The complexes prepared in this study were η6-diethyl(3-chloro-2-methyl) phenylmalonate- η5-cyclopentadienyliron hexafluorophosphate, η6-dimethyl(3-chloro-2-methyl)phenylmalonate-η5-cyclopentadienyliron hexafluorophosphate, η6-methyl(3-chloro-2-methyl)phenylacetoacetate-η5-cyclopentadienyliron hexafluorophosphate and η6-3-chloro(2-methyl-N-methyl)aniline-η5-cyclopentadienyliron hexafluorophosphate. Reaction of η6-2,6-dichlorotoluene-η5-cyclopentadienyliron hexafluorophosphate with excess methanol as well as methyl thioglycolate in the presence of K2CO3 resulted in disubstitution of both chloro groups to yield new complexes, η6-2,6-dimethoxytoluene-η5-cyclopentadienyliron hexafluorophosphate and η6-methyl[(2-methylphenyl)1,3-dithio] diacetate-η55-cyclopentadienyliron hexafluorophosphate, respectively. Complexes o-CH3C6H4CH(CO2C2H5)2FeCp+PF6, o-CH3C6H4CH(CO2CH3)2FeCp+PF6 and o-CH3C6H4CH2 COCH3FeCp+ PF6 react with excess K2CO3 and benzyl bromide in refluxing methylene chloride to give 80–90% yields of complexes o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6, o-CH3C6H4C(CH2C6H5)(CO2CH3)2FeCp+PF6 and o-CH3C6H4CH(CH2C6H5)COCH3FeCp+PF6, respectively. Reaction of complex, o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 with one molar equivalent of t-BuOK followed by acidic work-up gives o-(C2H5CO2CH2)C6H4CH(CO2C2H5)CH2C6H5FeCp+PF6. Similarly, reactions of complexes o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 and o-CH3C6H4C(CH2C6H5)(CO2CH3)2FeCp+PF6 with t-BuOK in THF followed by alkylation with methyl iodide gave the new complexes, o-(C2H5O2C(CH3)CH)C6H4CH(CH2C6H5)CO2C2H5FeCp+PF6 and o-(CH3O2C(CH3)CH)C6H4CH(CH2C6H5)CO2CH3FeCp+PF6, respectively. Vacuum sublimation of the new complexes, o-CH3C6H4C(CH2C6H5)(CO2C2H5)2FeCp+PF6 and o-(C2H5O2CCH2)C6H4CH(CH2C6H5)CO2C2H5FeCp+PF6 gives o-CH3C6H4C(CH2C6H5)(CO2C2H5)2 and O-(C2H5O2CCH2)C6H4CH(CH2C6H5)CO2C2H5, respectively.  相似文献   

17.
Cyclopentadienyl cobalt complexes (η5‐C5H4R) CoLI2 [L = CO,R=‐COOCH2CH=CH2 (3); L=PPh3, R=‐COOCH2‐CH=CH2 (6); L=P(p‐C6H4O3)3, R = ‐COOC(CH3) = CH2 (7), ‐COOCH2C6H5 (8), ‐COOCH2CH = CH2 (9)] were prepared and characterized by elemental analyses, 1H NMR, ER and UV‐vis spectra. The reaction of complexes (η5‐C5H4R)CoLI2 [L= CO, R= ‐COOC(CH3) = CH2 (1), ‐COOCH2C6H5(2); L=PPh3, R=‐COOC (CH3) = CH2 (4), ‐COOCH2C6H5 (5)] with Na‐Hg resulted in the formation of their corresponding substituted cobaltocene (η5‐C5H4R)2 Co[R=‐COOC(CH3) = CH2 (10), ‐COOCH2C6H5 (11)]. The electrochemical properties of these complexes 1–11 were studied by cyclic voltammetry. It was found that as the ligand (L) of the cobalt (III) complexes changing from CO to PPh3 and P(p‐tolyl)3, their oxidation potentials increased gradually. The cyclic voltammetry of α,α′‐substituted cobaltocene showed reversible oxidation of one electron process.  相似文献   

18.
Eight substituted bidentate Schiff base ligands HOC6H4CH=N-R (HL) (HL1: R = 4-ClC6H4, HL2: R = 2-ClC6H4, HL3: R = 4-NO2C6H4, HL4: R = 4-MeC6H4, HL5: R = 2,6-Me2C6H3, HL6: R = 2,46-Me3C6H2, HL7: R = CH2C6H5, and HL8: R = n-Pr) were synthesized by the typical condensation reaction. Interaction of cis-[Ru(bpy)2Cl2]?2H2O (bpy = 2,2′-bipyridine) with one equivalent of HL ligand in the presence of KPF6 afforded the cationic ruthenium(II) complexes of the type [Ru(bpy)2(L)](PF6) (18). The reaction of cis-[Ru(phen)2Cl2]?2H2O (phen = 1,10-phenanthroline) and HL1 under similar condition gave complex [(phen)2Ru(L)](PF6) (1a). Treatment of cis-[Ru(phen)2Cl2]?2H2O with two equivalents of HL in the presence of KPF6 resulted in isolation of the cationic ruthenium(III) complexes of the type [Ru(phen)(L)2](PF6) (9-16). All complexes have been spectroscopically characterized. The structures of 1a?CH2Cl2, 2?½CH2Cl2, 3?CH3CN, 5?½H2O, 6, 12?½HOCH2CH2OH, 13?CH3CN, 15?H2O, and 16 have been determined by single-crystal X-ray diffraction.  相似文献   

19.
The synthesis and characterization of optically active olefinic complexes of the type [(η-C5H5)Ru{Ph2PCH(CH3)CH2PPh2}(CH2CHR″)]PF6 (R″  H, CH3, C6H5, COOCH3), in which the metal is a stereogenic center, are reported. The enantioface discrimination of the prochiral olefin is influenced by the chiral ligand and by the stereogenic metal atom. The chiral center at the metal appears to be optically labile. The rates of the epimerization at the metal and of the olefin enantioface depend on the structure of the coordinated olefin.  相似文献   

20.
Imidazolium salts, [RS(O)? CH2(C3H3N2)Mes]Cl (R=Me ( L1 a ), Ph ( L1 b )); Mes=mesityl), make convenient carbene precursors. Palladation of L1 a affords the monodentate dinuclear complex, [(PdCl2{MeS(O)CH2(C3H2N2)Mes})2] ( 2 a ), which is converted into trans‐[PdCl2(NHC)2] (trans‐ 4 a ; N‐heterocyclic carbene) with two rotamers in anti and syn configurations. Complex trans‐ 4 a can isomerize into cis‐ 4 a (anti) at reflux in acetonitrile. Abstraction of chlorides from 4 a or 4 b leads to the formation of a new dication: trans‐[Pd{RS(O)CH2(C3H2N2)Mes}2](PF6)2 (R=Me ( 5 a ), Ph ( 5 b )). The X‐ray structure of 5 a provides evidence that the two bidentate SO? NHC ligands at palladium(II) are in square‐planar geometry. Two sulfoxides are sulfur‐ and oxygen‐bound, and constitute five‐ and six‐membered chelate rings with the metal center, respectively. In acetonitrile, complexes 5 a or 5 b spontaneously transform into cis‐[Pd(NHC)2(NCMe)2](PF6)2. Similar studies of thioether–NHCs have also been examined for comparison. The results indicate that sulfoxides are more labile than thioethers.  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