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.
Two new 1-D manganese(III) Schiff-base complexes bridged by dicyanamide (dca), [Mn(III)(5-Brsalen)(dca)] ? CH3OH (1) and [Mn(III)(3,5-Brsalen)(dca)] · CH3OH · CH3CN (2) (5-Brsalen = N,N′-ethylenebis(5-bromo salicylaldiminato) dianion; 3,5-Brsalen = N,N′-ethylenebis(3,5-dibromosalicylal diminato) dianion), have been synthesized and characterized. X-ray diffraction analyses reveal that the two complexes have 1-D chain structures constructed by μ 1,5-dca bridge. Magnetic susceptibility measurements exhibit weak antiferromagnetic exchange coupling in the complexes.  相似文献   

2.
Reaction of (rac)‐3,3′‐bis(methoxymethyl)‐BINOL [H2(CH3OCH2)2BINO] with excess Ti(OiPr)4 and one equivalent of H2O in CH2Cl2 affords a trinuclear titanium(IV) complex [{(CH3OCH2)2BINO}Ti3(μ3‐O)(OiPr)6(μ2‐OiPr)2]. By dissolving it in dichloromethane and hexane and cooling to 0 °C, plate‐like pale yellow single crystals (monoclinic, P21/n, a = 12.605(3), b = 21.994(5), c = 19.090(4) Å, β = 92.764(8)°, V = 5286.2(19) Å3, T = 293(2) K) were obtained. Each oxygen atom at 2 or 2′ position of the (CH3OCH2)2BINO ligand bonds to only one titanium atom. There is no interaction between the third Ti atom and the two oxygen atoms of 3,3′‐bis(methoxymethyl)‐BINOLate.  相似文献   

3.
Preparation and Properties of Trifluoromethylmercaptothiophosphoryldichloride The reaction of CF3SP(O)Cl2 with SPCl3 leads to a CF3S-chlorine exchange and gives CF3SP(S)Cl2 in 50% yield. A controlled hydrolysis of CF3SP(O)Cl2 affords CF3SP(O)(OH)2, that cannot be isolated as such, but it condenses to CF3SP(O)(OH)O? [P(SCF3)(O)? O]nP(O)(OH)SCF3. On the other hand, CF3SP(S)Cl2 reacts with water to yield H3PO4, CF3SH, S8, and HCl. CF3SP(X)Cl2 reacts with alcohols to give CF3SP(X)(OR)2 [R = CH3, C2H5, n-C3H7, CH(CH3)2, n-C4H9 and for X = O, R = C6H5, too]. The formation of semi-esters CF3SP(X)Cl(OR′) could be proven for X = O, R′ = CH3, C6H5 and for X = S, R′ = R. While CF3SP(O)(OC2H5)2 rapidly decomposes into SCF2 and FP(O)(OC2H5)2, the other compounds and primarily CF3SP(O)(OCH3)2 and CF3SP(S)(OR)2 ar stable. The reaction between CF3SCl and CH3SPCl2 results in CF3SCH2SPCl2 and that between CF3SP(O)Cl2 and AlCl3 gives [CF3SP(O)Cl]+[AlCl4]?. Physical and spectroscopical data are given for the newly formed compounds.  相似文献   

4.
Luminescent cuprous complexes are an important class of coordination compounds due to their relative abundance, low cost and ability to display excellent luminescence. The heteroleptic cuprous complex solvate rac‐(acetonitrile‐κN)(3‐aminopyridine‐κN)[2,2′‐bis(diphenylphosphanyl)‐1,1′‐binaphthyl‐κ2P,P′]copper(I) hexafluoridophosphate dichloromethane monosolvate, [Cu(C5H6N2)(C2H3N)(C44H32P2)]PF6·CH2Cl2, conventionally abbreviated as [Cu(3‐PyNH2)(CH3CN)(BINAP)]PF6·CH2Cl2, ( I ), where BINAP and 3‐PyNH2 represent 2,2′‐bis(diphenylphosphanyl)‐1,1′‐binaphthyl and 3‐aminopyridine, respectively, is described. In this complex solvate, the asymmetric unit consists of a cocrystallized dichloromethane molecule, a hexafluoridophosphate anion and a complete racemic heteroleptic cuprous complex cation in which the cuprous centre, in a tetrahedral CuP2N2 coordination, is coordinated by two P atoms from the BINAP ligand, one N atom from the 3‐PyNH2 ligand and another N atom from a coordinated acetonitrile molecule. The UV–Vis absorption and photoluminescence properties of this heteroleptic cuprous complex have been studied on polycrystalline powder samples, which had been verified by powder X‐ray diffraction before recording the spectra. Time‐dependent density functional theory (TD‐DFT) calculations and a wavefunction analysis reveal that the orange–yellow phosphorescence emission should originate from intra‐ligand (BINAP) charge transfer mixed with a little of the metal‐to‐ligand charge transfer 3(IL+ML)CT excited state.  相似文献   

5.
Five cationic complexes of the general formula [Cp′2Ti(A)2]2+ [Cl?]2 [Cp′ = η5‐(CH3)C5H4 and A = glycine, 1 ; 2‐methylalanine, 2 ; N‐methylglycine, 3 ; L ‐alanine, 4 ; and D ‐alanine 5 ] were prepared by the reaction of Cp′2TiCl2 and the appropriate α‐amino acid in 1:2 molar ratio from methanol–water solution in high yield. Air‐stable crystalline solids, highly soluble in water, were characterized by means of elemental analysis, IR, Raman, 1H, 13C and 14N NMR spectroscopy. The structure of compound 3 was determined by single crystal X‐ray crystallography: orthorhombic Pbca No. 61, a = 9.5310(3), b = 18.2980(5), c = 26.6350(5) Å, V = 4654 Å3, Z = 8. Hydrolytic stability of all compounds in D2O was investigated using 1H NMR spectroscopy within the pD interval of 2.9–6.5. All compounds slowly decomposed during 24 h at pD = 2.94, forming a mixture of hydrolytic products [Cp′2Ti(A)(D2O)]2+, [Cp′2Ti(D2O)2]2+ and respective α‐amino acids. By elevating pD to 4.0 and up to 6.5, a yellowish precipitate was formed, which indicates decomposition of the complexes. These compounds were characterized using elemental analyses, IR and Raman spectroscopy and attributed to oligomeric and/or polymeric structures described empirically by the formula Ti(Cp′)xOy(OH)z (x = 0.65; y = 0.3, z = 1.9). Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

6.
Supported Organometallic Complexes. VI. Characterization und Reactivity of Polysiloxane-Bound (Ether-phosphane)ruthenium(II) Complexes The ligands PhP(R)CH2D [R = (CH3O)3Si(CH2)3; D = CH2OCH3 ( 1b ); D = tetrahydrofuryl ( 1c ); D = 1,4-dioxanyl ( 1d )] have been used to synthesize (ether-phosphane)ruthenium(II) complexes, which have been copolymerized with Si(OEt)4 to yield polysiloxane-bound complexes. The monomers cis,cis,trans-Cl2Ru(CO)2(P ~ O)2 ( 3b ) and HRuCl(CO)(P ~ O)3 ( 5b ) were treated with NaBH4 to form cis,cis,trans-H2Ru(CO)2(P ~ O)2 ( 4b ) and H2Ru(CO)(P ~ O)3 ( 6b ), respectively (P ~ O = η1-P coordinated; = η2- coordinated). Addition of Si(OEt)4 and water leads to a base catalyzed hydrolysis of the silicon alkoxy-functions and a precipitation of the immobilized counterparts 4b ′, 6b ′. The polysiloxane matrix resulting by this new sol gel route has been described under quantitative aspects by 29Si CP-MAS NMR spectroscopy. 4b ′ reacts with carbon monoxide to form Ru(CO)3(P ~ O)2 ( 7b ′). Chelated polysiloxane-bound complexes Cl2Ru( )2 ( 9c ′, d ′) and Cl2Ru( )(P ~ O)2 ( 10b ′, c ′) have been synthesized by the reaction of 1b–c with Cl2Ru(PPh3)3 ( 8 ) followed by a copolymerization with Si(OEt)4. The polysiloxane-bound complexes 9c ′, d ′ and 10b ′, c ′ react with one equivalent of CO to give Cl2Ru(CO)( )(P ~ O) ( 12b ′– d ′). Excess CO leads to the all-trans-complexes Cl2Ru(CO)2(P ~ O)2 ( 14b ′– d ′), which are thermally isomerized to cis,cis,trans- 3b ′– d ′. The chemical shift anisotropy of 31P in crystalline Cl2Ru( )2 ( 9a , R = Ph, D = CH2OCH3) has been compared with polysiloxane-bound 9d ′ indicating a non-rigid behavior of the complexes in the matrix.  相似文献   

7.
A series of titanium(IV) complexes Ti(O‐i‐Pr)Cl3(THF)(PhCOR) (R = H ( 1 ), CH3 ( 2 ), or Ph ( 3 )) is prepared quantitatively from reactions of [Ti(O‐i‐Pr)Cl2(THF)(μ‐Cl)]2 with 2 molar equiv. PhCOR. Treatment of Ti(O‐i‐Pr)Cl3 with 2 molar equiv. of PhCOR affords the disubstituted complexes Ti(O‐i‐Pr)Cl3(PhCOR)2 (R = CH3 ( 4 ) or Ph ( 5 )). The 13C NMR study of these complexes shows that the relative bonding abilities are in the order of PhCOCH3 > PhCHO > PhCOPh. The molecular structure of 5 reveals that one of the benzophenone ligands is trans to the strongest 2‐propoxide ligand with a long Ti‐O(carbonyl) distance of 2.193(5) Å which is much longer than the other Ti‐O(carbonyl) distance of 2.097(4) Å by ?0.1 Å. All ligands cis to the alkoxide ligand are bending away from the alkoxide ligand with the RO‐Ti‐L angles ranging from 93.6(2) to 99.0(2)°.  相似文献   

8.
Reaction of [Au(DAPTA)(Cl)] with RaaiR’ in CH2Cl2 medium following ligand addition leads to [Au(DAPTA)(RaaiR’)](Cl) [DAPTA=diacetyl-1,3,5-triaza-7-phosphaadamantane, RaaiR’=p-R-C6H4-N=N- C3H2-NN-1-R’, (1—3), abbreviated as N,N’-chelator, where N(imidazole) and N(azo) represent N and N’, respectively; R=H (a), Me (b), Cl (c) and R’=Me (1), CH2CH3 (2), CH2Ph (3)]. The 1H NMR spectral measurements in D2O suggest methylene, CH2, in RaaiEt gives a complex AB type multiplet while in RaaiCH2Ph it shows AB type quartets. 13C NMR spectrum in D2O suggest the molecular skeleton. The 1H-1H COSY spectrum in D2O as well as contour peaks in the 1H-13C HMQC spectrum in D2O assign the solution structure.  相似文献   

9.
Luminescent cyclometalated platinum(II) complexes, namely [Pt(Thpy)(PPh3)X]n+ (HThpy = 2-(2′-thienyl)pyridine; X = Cl ( 1 ), n = 0; X = CH3CN ( 2 ), pyridine ( 3 ), n = 1) and [Pt(Thpy)(HThpy)Y] n + (Y = Cl ( 4 ), n = 0; Y = pyridine ( 5 ), n = 1), exhibit structured emission with peak maximum at ∼556 and 598 nm in degassed acetonitrile and with emission quantum yield and lifetime of up to 0.38 and 26 μs, respectively. These complexes are efficient photosensitizers of singlet oxygen with yields up to >90%. Complex 5 exhibited photocytotoxicity towards cancer cells and fluorescence microscopic images of cells incubated with 5 reveal substantial uptake at the nucleus and mitochondria.  相似文献   

10.
The reaction of C6Cl5Ni(PPhMe2)2Cl with CHRCR′CH2MgX (X = Br or Cl) yields π-allylnickel compounds, (π-CHRCR′CH2)Ni(PPhMe2)C6Cl5 (Ia, R = R′ = H; Ib, R = H, R′ = CH3; Ic, R = CH3, R′ = H), which are stable in the solid state below ca. 150°C and are fairly stable in solution in the absence of oxygen. The π-allyl group was found by PMR spectroscopy to be rigid even in the presence of an excess of PPhMe2, P(OEt)3 or P(OMe)3.  相似文献   

11.
α ω-Alkane-bis-dimethylarsine Sulfides and Selenides, a Novel Class of Ligands The reaction of α,ω-alkane-bis-dimethylarsanes (CH3)2As? (CH2)n? As (CH3)2 with sulfur and selenium results in formation of the sulfides and selenides, respectively, (CH3)2As(X)? (CH2)n? As(CH3)2 or (CH3)2As(X)? (CH2)n? As(X)(CH3)2 (X = S, Se), which form chelat-complexes with the salts CoX2 · 6 H2O (X = Cl?, Br?, I?, NO3?). The UV-spectra of the complexes are presented and discussed.  相似文献   

12.
Condensation polymerization of phosphonates through formation of P? O? P linkages has been achieved by (1) volatilization of methyl chloride from mixtures of CH3P(O)Cl2 with CH3P(O)(OCH3)2; (2) volatilization or chemical removal of water from CH3P(O)(OH)2; and (3) volatilization of HCl from mixtures of CH3P(O)Cl2 with CH3P(O)(OH)2 or C6H5P(O)Cl2 with C6H5P(O)(OH)2. Depending on the proportions of the reagents, the polymerization products consist of various mixtures of chain molecules of the type \documentclass{article}\pagestyle{empty}\begin{document}${\rm X \hbox{--} P}({\rm O})({\rm R})\rlap{--}[{\rm O \hbox{--} P}({\rm O})({\rm R})\rlap{--}]_n {\rm X}$\end{document} for R = CH3 and X = OCH3, Cl, or OH, or for R = C6H5, x = Cl or OH. 31P nuclear magnetic resonance (NMR) was used to investigate both the polymethylpolyphosphonates and the polyphenylpolyphosphonates; and 1H NMR of the CH3P and CH3O moieties was also used to study the polymethylpolyphosphonates. In the methoxyl-terminated polymethylpolyphosphonates, which was the system studied most extensively, no detectable amounts of cyclic molecules were found at equilibrium, but a crystalline methylphosphonic anhydride, CH3PO2, exhibited some ring structures. The equilibrium size distributions gave evidence that the sorting of the mono- and difunctional phosphorus-based units making up the oligomeric chains is affected by neighboring units. Kinetic measurements demonstrated that the condensation polymerization is a complicated process involving considerable scrambling of terminal groups with bridging oxygen atoms.  相似文献   

13.
Bulk free radical polymerization of the monomer series CH2 = C(CH3)C(O)OCH2CH3‐n Cln , n = 1, 2, 3, yields an unexpectedly crosslinked product with a crosslink density that increases with decreasing chlorine content of the respective monomer (n = 3 < n = 2 < n = 1). This chlorine substituent effect is investigated by correlation with chain transfer constant measurements for four homologous series of chloroalkyl compounds (chloroethyl acetates (CH3C(O)OCH2CH3‐n Cln , n = 1,2,3); chloromethanes (CH4‐n Cln , n = 2,3,4) and CD2Cl2 and CDCl3 analogs; butyl chloride isomers (n‐ , iso‐ , sec‐, tert‐) and tert‐C4D9Cl analog; and nine chloroethanes (C2Hn ?6Cln , n = 1–6)) in a methyl methacrylate polymerization. The pattern conveyed by the magnitude of chain transfer constants and deuterium isotope effects is consistent with a vicinal chlorine effect (i.e., chlorine activation of a vicinal hydrogen for abstraction) to account for the relative activities of the four series of model compounds and for the propensity of the chloroethyl methacrylates to crosslink in a bulk free radical polymerization. The chloroalkyl moiety's contribution to chain transfer is relatively modest (≤10?4), but, when incorporated as a monomer pendant group in free radical polymerizations, it is effective in broadening molecular weight to the extent of resulting in a crosslinked polymer. Published 2016.? J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 93–106  相似文献   

14.
The complex formation of PdII with tris[2-(dimethylamino)ethyl]amine (N(CH2CH2N(CH3)2)3, Me6tren) was investigated at 25° and ionic strength I = 1, using UV/VIS, potentiometric, and NMR measurements. Chloride, bromide, and thiocyanate were used as auxiliary ligands. The stability constant of [Pd(Me6tren)]2+ in various ionic media was obtained: log β([Pd(Me6tren)] = 30.5 (I = 1(NaCl)) and 30.8 (I = 1(NaBr)), as well as the formation constants of the mixed complexes [Pd(HMe6tren)X]2+ from [Pd(HMe6tren)(H2O)]3+:log K = 3.50 = Cl?) and 3.64 (X? = Br?) and [Pd(Me6tren)X]+ from [Pd(Me6tren)(H2O)]2+: log K = 2.6 (X? = Cl?), 2.8(Br?) and 5.57 (SCN?) at I = 1 (NaClO3). The above data, as well as the NMR measurements do not provide any evidence for the penta-coordination of PdII, proposed in some papers.  相似文献   

15.
The solution properties of a series of transition‐metal–ligand coordination polymers [ML(X)n] [M=AgI, ZnII, HgII and CdII; L=4,4′‐bipyridine (4,4′‐bipy), pyrazine (pyz), 3,4′‐bipyridine (3,4′‐bipy), 4‐(10‐(pyridin‐4‐yl)anthracen‐9‐yl)pyridine (anbp); X=NO3?, CH3COO?, CF3SO3?, Cl?, BF4?; n=1 or 2] in the presence of competing anions, metal cations and ligands have been investigated systematically. Providing that the solubility of the starting complex is sufficiently high, all the components of the coordination polymer, namely the anion, the cation and the ligand, can be exchanged on contact with a solution phase of a competing component. The solubility of coordination polymers is a key factor in the analysis of their reactivity and this solubility depends strongly on the physical properties of the solvent and on its ability to bind metal cations constituting the backbone of the coordination polymer. The degree of reversibility of these solvent‐induced anion‐exchange transformations is determined by the ratio of the solubility product constants for the starting and resultant complexes, which in turn depend upon the choice of solvent and the temperature. The extent of anion exchange is controlled effectively by the ratio of the concentrations of incoming ions to outgoing ions in the liquid phase and the solvation of various constituent components comprising the coordination polymer. These observations can be rationalised in terms of a dynamic equilibrium of ion exchange reactions coupled with Ostwald ripening of crystalline products. The single‐crystal X‐ray structures of [Ag(pyz)ClO4] ( 1 ), {[Ag(4,4′‐bipy)(CF3SO3)] ? CH3CN} ( 2 ), {[Ag(4,4′‐bipy)(CH3CN)]ClO4 ? 0.5 CH3CN} ( 3 ), metal‐free anbp ( 4 ), [Ag(anbp)NO3(H2O)] ( 5 ), {[Cd(4,4′‐bipy)2(H2O)2](NO3)2 ? 4 H2O} ( 6 ) and {[Zn(4,4′‐bipy)SO4(H2O)3] ? 2 H2O} ( 7 ) are reported.  相似文献   

16.
Electrochemical investigations of the reduction of dicationic, monocationic and neutral dinitrosyl molybdenum complexes in nitromethane and acetonitrile are reported. All the compounds with the general formulae: [Mo(NO)2L2L′2]2+, [Mo(NO)2L2L′Cl]+ and Mo(NO)2L2Cl2 (L = CH3CN, CH2CHCN, C6H5CN, C5H5N, P(C6H5)3, L2 = 2,2′-bipyridine, L′ = CH3CN and L′2 = 2,2′-bipyridine) are reducible by one electron to yield 19-electron complexes. The dicationic complexes undergo a reversible one-electron transfer. For the mono- and dichlorocomplexes, the one-electron transfer induces the facile exchange of the chloroligand in the 19-electron complexes except for L2 = 2,2′-bipyridine. However, the exchange of the chloroligand is followed by the fast anation by Cl? of the remaining 18-electron chlorocomplexes to afford [Mo(NO)2Cl3L]? and [Mo(NO)2Cl4]2? which are reducible at higher negative potentials than dichloro- and monochlorocomplexes. The multiple electrochemical step system is not catalytic, but of the electroactivation type.  相似文献   

17.
The reaction between Cl2Te(NSO)2, Cl6Te2N2S and Cl2Te(N=S=N)2TeCl2 with MCl3 provided the compounds [(Cl2Te)2N+][MCl4] (M = Ga, Al, Fe). Treating Cl6Te2N2S with M′Cl3 yielded besides [(Cl2Te)2N+][M′Cl4] (M′ = Al, Fe) the sulfur containing compound [ClTeNSNS+][M′Cl4]. The structure for [ClTeNSNS+][FeCl4] was established by an X‐ray structure analysis. With Te(NSO)2 and CF3SCl, via Cl2Te(NSO)2, the known compound Te2NCl5 was formed. Tetrafluoroditelluradiazetidine was obtained from TeF4 and [(CH3)3Si]2NH which on treating with (CH3)3SiCl provided the corresponding chloroderivative. In addition metathetical reaction between Cl2TeNSNS and CF3C(O)OAg yielded [CF3C(O)O]2TeSNSN. Similarly (CH3)2Te(NSO)2–xClx (x = 0,1) and (CH3)2Te(NCO)2 were made from (CH3)2TeCl2 and AgNSO or AgNCO, respectively. Halogination of Cl2Te(N=S=N)2TeCl2 with Cl2 or Br2 yielded Cl6Te2N2S and Cl4Br2Te2N2S. The bromoderivate was also prepared from Cl2Te(NSO)2 and Br2. AgNSO was synthesized by treating CF3C(O)OAg with (CH3)3SiNSO. Two other synthons (CF3Se)2Te and (CF3S)2Se were obtained from CF3SeCl and Na2Te and from Hg(SCF3)2 plus SeCl4, respectively.  相似文献   

18.
The geometric structure of the ground state and of metastable isomers of nitrosyl complexes trans-[Ru(P)(NO)(Cl)] (P = porphinate dianion) and trans-[Ru(NO)(salen)(X)]q [salen = N,N'-ethylenebis(salicylideniminate) dianion; X = Cl- (q = 0), H2O (q = +1)] was optimized within the framework of the density functional method (SVWN/LanL2DZ+6-31G). The local minima corresponding to metastable isomers with a linear NO coordination through the oxygen atom and with a side 2 NO coordination were found on the potential energy surfaces of these compounds. The second metastable states of all the three complexes have a lower energy. The difference in energies between the stable and metastable isomers is the least in the case of the complex trans-[Ru(NO)(salen)(Cl)].  相似文献   

19.
Four heterodimetallic complexes [Ru(Fcdpb)(L)](PF6) (Fcdpb=2‐deprotonated form of 1,3‐di(2‐pyridyl)‐5‐ferrocenylbenzene; L=2,6‐bis‐(N‐methylbenzimidazolyl)‐pyridine (Mebip), 2,2′:6′,2′′‐terpyridine (tpy), 4‐nitro‐2,2′:6′,2′′‐terpyridine (NO2tpy), and trimethyl‐4,4′,4′′‐tricarboxylate‐2,2′:6′,2′′‐terpyridine (Me3tctpy)) have been prepared. The electrochemical and spectroelectrochemical properties of these complexes have been examined in CH2Cl2, CH3NO2, CH3CN, and acetone. These complexes display two consecutive redox couples owing to the stepwise oxidation of the ferrocene (Fc) and ruthenium units, respectively. The potential difference, ΔE1/2 (E1/2(RuII/III)?E1/2(Fc0/+)), decreased slightly with increasing solvent donocity. The mixed‐valent states of these complexes have been generated by electrolysis and the resulting intervalence charge‐transfer (IVCT) bands have been analyzed by Hush theory. Good linear relationships exist between the energy of the IVCT band, Eop, and ΔE1/2 of four mixed‐valent complexes in a given solvent.  相似文献   

20.
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   

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