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 515 毫秒
1.
4-Hydroxyacridine (HAcr) is an O,N-chelating ligand whose coordination chemistry toward group 13 M(III) ions has received little attention. The molecular structure of HAcr consists of a 2,3-disubstituted-8-hydroxyquinoline; thus, in order to compare 8-hydroxyquinoline (HQ), 2-methyl-8-hydroxyquinoline (HMeQ′), and 2,3-disubstituted-8-hydroxyquinoline (HAcr) for steric and/or electronic influence, HAcr chelating ability toward the Al(III), Ga(III), and In(III) triad has been investigated. Irrespective of the nature of M(III), only complexes containing two equivalents of deprotonated HAcr are obtained. This article describes the synthesis and characterization of different series of bis-chelated pentacoordinated (Acr)2MY (M = Al, Ga, In; Y = Cl, Br, I, NCS, N3) or (Acr)2MZ (M = Ga or In; HZ = C6H5OH, C6H13OC6H4OH, C6H5COOH, or C6H13OC6H4COOH) six-coordinate neutral (Acr)2In(acac) (H(acac) =acetylacetone), or ionic [(Acr)2In(N,N)][CF3SO3] (N,N = 2,2′-bipyridine or 1,10-phenanthroline) complexes. These results significantly contribute to elucidating the complexation capability of HAcr.  相似文献   

2.
《Polyhedron》1988,7(14):1289-1298
The following adducts of Group III trialkyls with phosphines have been prepared, either by direct reaction in hydrocarbon solution or by displacement of ether from the metal trialkyl etherate: Me3M·PPh3 (M = Ga, In); Me3In·P(2-MeC6H4)3; (R3M)2·(Ph2PCH2)2 (R = Me, M = Al, Ga, In; R = Et, M = Ga, In; R = Bui, M = Al); (Me3M)3·(Ph2PCH2CH2)2PPh (M = Al, Ga, In) and (Me3M)4·(Ph2PCH2CH2PPhCH2)2 (M = Al, Ga, In). The compounds were analysed by 1H and 31P NMR spectra of (Me3M)2·(Ph2PCH2)2 (M = Ga, In) showed little change between 193 K and room temperature. Thermal dissociation of the adducts in vacuo gave the free metal trialkyl with no detectable contamination by the respective phosphine. Crystals of (Me3M)2·(Ph2PCH2)2 (M = Al, Ga, In) are iso-structural and the molecules contain two distorted tetrahedral metals bridged by the (Ph2PCH2)2; the MP distances are 2.544(4), 2.546(4) and 2.755(4) Å, respectively. The X-ray crystal structure of (Me3Al)3·(Ph2PCH2CH2)2PPh shows the molecule to contain distorted tetrahedral aluminium atoms bonded to each of the three phosphorus atoms, with AlP distances of 2.536(9) and 2.510(9) Å for the terminal and central moieties, respectively; the unit cell contains two such molecules plus one benzene molecule (the crystallizing solvent).  相似文献   

3.
Perfluoromethyl Element Ligands. XLIII [1] Novel Synthetic Routes to Binuclear Complexes of the Type MM′(CO)8ER2X (M/M′ = Mn/Mn, Mn/Re, Re/Re; E = P, As; R = CF3, Me; X = Hal, ) Mn(CO)5I reacts with compounds of the type (CF3)2EAsMe2 (E = P, As) as with the symmetric E2(CF3)4 ligands in the first step with cleavage of the E‐As bond to yield the pro ducts (CO)5MnE(CF3)2 and Me2AsI. Reaction of the mononuclear complexes with excess of Mn(CO)5I leads in good yields to the known dinuclear compounds (CO)4Mn[E(CF3)2, I]Mn(CO)4 and CO. Me2AsI, the second product of the EAs cleavage, attacks the starting compound Mn(CO)5I giving cis‐Mn(CO)4I(AsMe2I) and CO. This result encouraged us to thoroughly investigate the preparation of cis‐M(CO)4X(EMe2Y) complexes with most of the possible combinations of M = Mn, Re; E = P, As and X, Y = Cl, Br, I. An alternative route to these compounds was opened by the cleavage of the dinuclear manganese or rhenium halides M2(CO)8X2 with the halophosphanes or ‐arsanes Me2EY. This route was found to be especially advantageous for the preparation of the rheniumcarbonyl precursors, since milder conditions than for the CO‐substitution in Re(CO)5X compounds are sufficient for the halogen‐bridged dinuclear complexes. Cis‐M(CO)4X(EMe2Y) complexes were used as precursors for the synthesis of novel homo‐ and heterodinuclear complexes of the type (CO)4M(EMe2, X)M′(CO)4 by reacting the EY function with transition metal carbonylates Kat[M′(CO)5] (Kat = Na, Bu4N, Ph4As). Thus the preparation of a wide range of complexes was possible, which before had been successfully prepared by the direct reaction of Mn2(CO)10 with Me2EX only in few cases, e. g. with Me2AsI. Spectroscopic investigations, using the CO valence frequencies and the 1H‐NMR data of the ligands EMe2Y or of the Me2E bridges, were applied to study the influence of the variables M, M′, E, X, Y and Kat on the reactivity of the mononuclear complexes and the bonding situation in both the mono‐ and the dinuclear systems. The new compounds were characterized by spectroscopic (IR, NMR, MS) and analytic methods (C, H).  相似文献   

4.
Chloride abstraction from the complexes [(η6-p-cymene){(IDipp)P}MCl] ( 2 a , M=Ru; 2 b , M=Os) and [(η5-C5Me5){(IDipp)P}IrCl] ( 3 b , IDipp=1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene) with sodium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate (NaBArF) in the presence of trimethylphosphine (PMe3), 1,3,4,5-tetramethylimidazolin-2-ylidene (MeIMe) or carbon monoxide (CO) afforded the complexes [(η6-p-cymene){(IDipp)P}M(PMe3)]BArF] ( 4 a , M=Ru; 4 b , M=Os), [(η6-p-cymene){(IDipp)P}Os(MeIMe)]BArF] ( 5 ) and [(η5-C5Me5){(IDipp)P}IrL][BArF] ( 6 , L=PMe3; 7 , L=MeIMe; 8 , L=CO). These cationic N-heterocyclic carbene-phosphinidene complexes feature very similar structural and spectroscopic properties as prototypic nucleophilic arylphosphinidene complexes such as low-field 31P NMR resonances and short metal-phosphorus double bonds. Density functional theory (DFT) calculations reveal that the metal-phosphorus bond can be described in terms of an interaction between a triplet [(IDipp)P]+ cation and a triplet metal complex fragment ligand with highly covalent σ- and π-contributions. Crystals of the C−H activated complex 9 were isolated from solutions containing the PMe3 complex, and its formation can be rationalized by PMe3 dissociation and formation of a putative 16-electron intermediate [(η5-C5Me5)Ir{P(IDipp)}I][BArF], which undergoes C−H activation at one of the Dipp isopropyl groups and addition along the iridium-phosphorus bond to afford an unusual η3-benzyl coordination mode.  相似文献   

5.
The reactions of [Ru(N2)(PR3)(‘N2Me2S2’)] [‘N2Me2S2’=1,2‐ethanediamine‐N,N′‐dimethyl‐N,N′‐bis(2‐benzenethiolate)(2?)] [ 1 a (R=iPr), 1 b (R=Cy)] and [μ‐N2{Ru(N2)(PiPr3)(‘N2Me2S2’)}2] ( 1 c ) with H2, NaBH4, and NBu4BH4, intended to reduce the N2 ligands, led to substitution of N2 and formation of the new complexes [Ru(H2)(PR3)(‘N2Me2S2’)] [ 2 a (R=iPr), 2 b (R=Cy)], [Ru(BH3)(PR3)(‘N2Me2S2’)] [ 3 a (R=iPr), 3 b (R=Cy)], and [Ru(H)(PR3)(‘N2Me2S2’)]? [ 4 a (R=iPr), 4 b (R=Cy)]. The BH3 and hydride complexes 3 a , 3 b , 4 a , and 4 b were obtained subsequently by rational synthesis from 1 a or 1 b and BH3?THF or LiBEt3H. The primary step in all reactions probably is the dissociation of N2 from the N2 complexes to give coordinatively unsaturated [Ru(PR3)(‘N2Me2S2’)] fragments that add H2, BH4?, BH3, or H?. All complexes were completely characterized by elemental analysis and common spectroscopic methods. The molecular structures of [Ru(H2)(PR3)(‘N2Me2S2’)] [ 2 a (R=iPr), 2 b (R=Cy)], [Ru(BH3)(PiPr3)(‘N2Me2S2’)] ( 3 a ), [Li(THF)2][Ru(H)(PiPr3)(‘N2Me2S2’)] ([Li(THF)2]‐ 4 a ), and NBu4[Ru(H)(PCy3)(‘N2Me2S2’)] (NBu4‐ 4 b ) were determined by X‐ray crystal structure analysis. Measurements of the NMR relaxation time T1 corroborated the η2 bonding mode of the H2 ligands in 2 a (T1=35 ms) and 2 b (T1=21 ms). The H,D coupling constants of the analogous HD complexes HD‐ 2 a (1J(H,D)=26.0 Hz) and HD‐ 2 b (1J(H,D)=25.9 Hz) enabled calculation of the H? D distances, which agreed with the values found by X‐ray crystal structure analysis ( 2 a : 92 pm (X‐ray) versus 98 pm (calculated), 2 b : 99 versus 98 pm). The BH3 entities in 3 a and 3 b bind to one thiolate donor of the [Ru(PR3)(‘N2Me2S2’)] fragment and through a B‐H‐Ru bond to the Ru center. The hydride complex anions 4 a and 4 b are extremely Brønsted basic and are instantanously protonated to give the η2‐H2 complexes 2 a and 2 b .  相似文献   

6.
The limits of steric crowding in organometallic metallocene complexes have been examined by studying the synthesis of [(C5Me5)3MLn] complexes as a function of metal in which L=Me3CCN, Me3CNC, and Me3SiCN. The bis(tert‐butyl nitrile) complexes [(C5Me5)3Ln(NCCMe3)2] (Ln=La, 1 ; Ce, 2 ; Pr, 3 ) can be isolated with the largest lanthanide metal ions, La3+, Ce3+, and Pr3+. The Pr3+ ion also forms an isolable mono‐nitrile complex, [(C5Me5)3Pr(NCCMe3)] ( 4 ), whereas for Nd3+ only the mono‐adduct [(C5Me5)3Nd(NCCMe3)] ( 5 ) was observed. With smaller metal ions, Sm3+ and Y3+, insertion of Me3CCN into the M? C(C5Me5) bond was observed to form the cyclopentadiene‐substituted ketimide complexes [(C5Me5)2Ln{NC(C5Me5)(CMe3)}(NCCMe3)] (Ln=Sm, 6 ; Y, 7 ). With tert‐butyl isocyanide ligands, a bis‐isocyanide product can be isolated with lanthanum, [(C5Me5)3La(CNCMe3)2] ( 8 ), and a mono‐isocyanide product with neodymium, [(C5Me5)3Nd(CNCMe3)] ( 9 ). Silicon–carbon bond cleavage was observed in reactions between [(C5Me5)3Ln] complexes and trimethylsilyl cyanide, Me3SiCN, to produce the trimeric cyanide complexes [{(C5Me5)2Ln(μ‐CN)(NCSiMe3)}3] (Ln=La, 10 ; Pr, 11 ). With uranium, a mono‐nitrile reaction product, [(C5Me5)3U(NCCMe3)] ( 12 ), which is analogous to 5 , was obtained from the reaction between [(C5Me5)3U] and Me3CCN, but [(C5Me5)3U] reacts with Me3CNC through C? N bond cleavage to form a trimeric cyanide complex, [{(C5Me5)2U(μ‐CN)(CNCMe3)}3] ( 13 ).  相似文献   

7.
The preparation and characterization of a series of neutral rare‐earth metal complexes [Ln(Me3TACD)(η3‐C3H5)2] (Ln=Y, La, Ce, Pr, Nd, Sm) supported by the 1,4,7‐trimethyl‐1,4,7,10‐tetraazacyclododecane anion (Me3TACD?) are reported. Upon treatment of the neutral allyl complexes [Ln(Me3TACD)(η3‐C3H5)2] with Brønsted acids, monocationic allyl complexes [Ln(Me3TACD)(η3‐C3H5)(thf)2][B(C6X5)4] (Ln=La, Ce, Nd, X=H, F) were isolated and characterized. Hydrogenolysis gave the hydride complexes [Ln(Me3TACD)H2]n (Ln=Y, n=3; La, n=4; Sm). X‐ray crystallography showed the lanthanum hydride to be tetranuclear. Reactivity studies of [Ln(Me3TACD)R2]n (R=η3‐C3H5, n=0; R=H, n=3,4) towards furan derivatives includes hydrosilylation and deoxygenation under ring‐opening conditions.  相似文献   

8.
The antimony aminoalkoxide and aminothiolates Sb(ECH2CH2NMe2)3 [E = O ( 1 ), S ( 2 )] were synthesized and their ability to form adducts with other metal moieties investigated. Compound 1 forms 1:1 adducts with NiI2 ( 3 ) and M(acac)2 [M = Cd ( 4 ), Ni ( 5 )], while 2 undergoes ligand exchange with AlMe3 to afford Me2AlSCH2CH2NMe2 ( 6 ). The structures of 2 – 4 and 6 were determined. Compound 2 incorporates three S, N‐chelating ligands though the interaction with nitrogen is weaker than in analogous alkoxide complexes. Product 3 reveals one iodine has migrated from nickel to antimony, and all three alkoxide ligands bridge the two metals through μ2‐O atoms. In contrast, in 4 , only one alkoxide links the antimony and cadmium. Compound 6 adopts the same structure, a chelating S,N ligand generating a tetrahedral center at aluminum, as known tBu2AlSCH2CH2NR2 species (R = Me, Et).  相似文献   

9.
Substitution Reactions of Bis(trimethylelement)carbodiimides of Silicon and Germanium with Metal Chlorides and Dimethylmetalchlorides of Sb, Al, Ga, and In The reaction of Me3Ge? N?C?N? GeMe3 (Me?CH3) with SbCl5 in a 1 : 1 molar ratio forms dimeric Cl4SbNCNGeMe3 in high yields. The corresponding compounds (X2MNCNSiMe3)2–3 (with X?Cl, Me and M = Al, Ga), formed by reactions of X2MCl and Me3SiNCNSiMe3, are less stable and tend to condensations, eliminating Me3SiX. The carbodiimide derivates (Me2MNCNEMe3)2–3 (with E = Si, Ge) are also available in aprotic solvents from polymeric LiNCNEMe3 and Me2MCl (M = Al, Ga, In). According to the IR and Raman spectra the low associated substitution products consists of cyclic ring skeletons and asymmetric > N? C?N? EMe3 units with cyanamide conformation.  相似文献   

10.
Reaction of (R,R)‐(N,N′)‐Diisopropylcyclohexyl‐1,2‐diamine with Me2MCl (M = Ga, In) (R,R)‐(N,N′)‐Diisopropylcyclohexyl‐1,2‐diamine (H2L) was reacted with Me2GaCl and Me2InCl in boiling toluene, respectively. In both cases the salt [Me2M(H2L)][Me2MCl2] [M = Ga ( 1 ), In ( 2 )] was formed. 1 and 2 were characterized by NMR and vibrational spectroscopy. In addition, an X‐ray structure determination was applied on 2 . According to the spectroscopical and structural findings 1 and 2 consist of cations [Me2M(H2L)]+ and anions [Me2MCl2]?.  相似文献   

11.
The reaction of N-phthaloylglycine and N-phthaloyl-dl-alanine with trimethylgallium (1:1) yielded the dinuclear complexes [Me2Ga(μ-O2CCH2N(CO)2C6H4)]2 (1) and RS-[Me2Ga(μ-O2CCHMeN(CO)2C6H4)]2 (2), respectively. The molecular structure of 2 was determined by X-ray diffraction studies. The cytotoxic activity of the organogallium(III) complexes (1 and 2) was tested against human tumour cell lines 8505C anaplastic thyroid cancer, A253 head and neck tumour, A549 lung carcinoma, A2780 ovarian cancer, DLD-1 colon carcinoma and compared with that of cisplatin.The best response of the synthesized gallium complexes, compared with that of cisplatin, was observed against 8505C anaplastic thyroid cancer and DLD-1 colon carcinoma, while the best IC50 values were found for A253 head and neck carcinoma. While the studied carboxylic acids show no proliferative activity, complexes 1 and 2 present very similar cytotoxic activity against all the studied cancer cell lines (IC50 from ca. 5 to 25 μM). The cytotoxicities of complexes 1 and 2 are in all cases higher than that presented by gallium(III) nitrate. In addition DNA laddering method showed that treatment of the studied cell lines with IC90 doses of 1 and 2 resulted in the induction of apoptotic mode of cell death.  相似文献   

12.
Reaction of the binuclear μ‐carbamoyl complex [(CO)3Fe(μ‐Me2NCO)2Fe(CO)2(HNMe2)] ( 1 ) in toluene with the chelating ligands Ph2PCH2PPh2 (dppm) and Ph2PCH2CH2PPh2 (dppe) gives different results. With dppm only the complex [(CO)3Fe(μ‐Me2NCO)2Fe(CO)2(dppm)] ( 3 ) with a dangling ligand is obtained under replacement of amine, whereas with dppe depending on the reaction conditions up to three compounds are found. A 1 : 1 mixture of the educts generates the related complex [(CO)3Fe(μ‐Me2NCO)2Fe(CO)2(dppe)] ( 4 ) together with the tetranuclear complex [{(CO)3Fe(μ‐Me2NCO)2Fe(CO)2}2(dppe)] (5 ). 4 slowly converts into [(CO)3Fe(μ‐Me2NCO)2Fe(CO)(dppe)] ( 6 ) with dppe acting as a chelating ligand. 6 is the first compound in this series in which one of the five CO groups is replaced by another donor. A 2 : 1 molar ratio of 1 and dppe quantitatively produces 5 . Addition of CO to a solution of 6 proceeds under slow reversible conversion of the complex into 4 . The compounds were characterized by the usual spectroscopic methods; 3 , 5 and 6 were also studied by X‐ray diffraction analyses.  相似文献   

13.
Utilization of the N,C,N‐chelating ligand L (L={2,6‐(Me2NCH2)2C6H3}?) in the chemistry of 13 group elements provided either N→In coordinated monomeric chalcogenides LIn(μ‐E4) (E=S, Se) with unprecedented InE4 inorganic ring or monomeric chalcogenolates LM(EPh)2 (M=Ga, In). Complex LGa(SePh)2 was selected as the most suitable single source precursor (SSP) for the deposition of amorphous semiconducting GaSe thin films using spin coating method.  相似文献   

14.
A series of M(II) and M(IV) (M=Mo, W) alkyne adducts employing two 6-methylpyridine-2-thiolate (6-MePyS) ligands was synthesized and investigated towards the nucleophilic attack of PMe3 on the coordinated alkynes. For this approach, 2-butyne (C2Me2), phenylacetylene (HC2Ph), and diphenylacetylene (C2Ph2) were used. For the exploration of an intramolecular attack, but-3-yn-1-ol (HCCCH2CH2OH) was coordinated to the metal centers. A nucleophilic attack of PMe3 was observed in [W(CO)(HC2Ph)(6-MePyS)2] yielding an η2-vinyl compound. Reaction of [W(CO)(C2Ph2)(6-MePyS)2] with excess PMe3 resulted in the selective coordination of one molecule of PMe3 concomitant with decoordination of the nitrogen atom of one 6-MePyS ligand. In contrast, the W(IV) complexes did not react with PMe3. While no selectivity was observed in the reaction of the Mo(II) compounds with PMe3, alkynes in the Mo(IV) compounds were replaced by PMe3. Addition of Et3N to the but-3-yn-1-ol complexes did not lead to the anticipated formation of 2,3-dihydrofuran.  相似文献   

15.
Reactions of Group 4 metallocene alkyne complexes [Cp′2M(η2‐Me3SiC2SiMe3)] ( 1 : M=Zr, Cp′=Cp*=η5‐pentamethylcyclopentadienyl; 2 a : M=Ti, Cp′=Cp*, and 2 b : M=Ti, Cp′2=rac‐(ebthi)=rac‐1,2‐ethylene‐1,1′‐bis(η5‐tetrahydroindenyl)) with diphenylacetonitrile (Ph2CHCN) and of the seven‐membered zirconacyclocumulene 3 with phenylacetonitrile (PhCH2CN) were investigated. Different compounds were obtained depending on the metal, the cyclopentadienyl ligand and the reaction temperature. In the first step, Ph2CHCN coordinated to 1 to form [Cp*2Zr(η2‐Me3SiC2SiMe3)(NCCHPh2)] ( 4 ). Higher temperatures led to elimination of the alkyne, coordination of a second Ph2CHCN and transformation of the nitriles to a keteniminate and an imine ligand in [Cp*2Zr(NC2Ph2)(NCHCHPh2)] ( 5 ). The conversion of 4 to 5 was monitored by using 1H NMR spectroscopy. The analogue titanocene complex 2 a eliminated the alkyne first, which led directly to [Cp*2Ti(NC2Ph2)2] ( 6 ) with two keteniminate ligands. In contrast, the reaction of 2 b with diphenylacetonitrile involved a formal coupling of the nitriles to obtain the unusual four‐membered titanacycle 7 . An unexpected six‐membered fused zirconaheterocycle ( 8 ) resulted from the reaction of 3 with PhCH2CN. The molecular structures of complexes 4 , 5 , 6 , 7 and 8 were determined by X‐ray crystallography.  相似文献   

16.
A one pot reaction of Li2{1, 4‐(Me3Si)2C8H6}, LnCl3, and K{CH(PPh2NSiMe3)2} leads to the 1, 4‐bis(trimethylsilyl)cyclooctatetraene bis(phosphinimino)methanide complexes of yttrium and erbium, [{CH(PPh2NSiMe3)2}Ln(η8‐{1, 4‐(Me3Si)2C8H6})] (Ln = Y, Er). Both complexes have been characterized by single crystal X‐ray diffraction. The solid state structures show that the two bulky ligands cause a steric crowding around the lanthanide atom. As a result of this steric crowding both ligands are asymmetrically attached to the lanthanide atom.  相似文献   

17.
18.
The reaction of [(η5-C5Me5)M(μCl)Cl]2 with the ligand (LL) in the presence of sodium methoxide yielded compounds of general formula [(η5-C5Me5)M(LL)Cl] (1–10) (where M = Ir or Rh and LL = NO or OO chelate ligands). Azido complexes of formulation [(η5-C5Me5)M(LL)N3] (11–20) have been prepared by the reaction of [(η5-C5Me5)M(μN3)(X)]2 (X = Cl or N3) with the corresponding ligands or by the direct reaction of [(η5-C5Me5)M(LL)Cl] with NaN3. These azido complexes [(η5-C5Me5)M(LL)N3] undergo 1,3-dipolar cycloaddition reaction with substituted alkynes in CH2Cl2 and for the first time in ethanol at room temperature to yield iridium (III) and rhodium (III) triazoles (21–28). The compounds were characterized on the basis of spectroscopic data, and the molecular structures of 2 and 26 have been established by single crystal X-ray diffraction.  相似文献   

19.
The reaction of fac‐[MIIIF3(Me3tacn)]?x H2O with Gd(NO3)3?5H2O affords a series of fluoride‐bridged, trigonal bipyramidal {GdIII3MIII2} (M=Cr ( 1 ), Fe ( 2 ), Ga ( 3 )) complexes without signs of concomitant GdF3 formation, thereby demonstrating the applicability even of labile fluoride‐complexes as precursors for 3d–4f systems. Molecular geometry enforces weak exchange interactions, which is rationalized computationally. This, in conjunction with a lightweight ligand sphere, gives rise to large magnetic entropy changes of 38.3 J kg?1 K?1 ( 1 ) and 33.1 J kg?1 K?1 ( 2 ) for the field change 7 T→0 T. Interestingly, the entropy change, and the magnetocaloric effect, are smaller in 2 than in 1 despite the larger spin ground state of the former secured by intramolecular Fe–Gd ferromagnetic interactions. This observation underlines the necessity of controlling not only the ground state but also close‐lying excited states for successful design of molecular refrigerants.  相似文献   

20.
New iron complexes [Cp*Fe L ]? ( 1‐σ and 1‐π , Cp*=C5Me5) containing the chelating phosphinine ligand 2‐(2′‐pyridyl)‐4,6‐diphenylphosphinine ( L ) have been prepared, and found to undergo facile reaction with CO2 under ambient conditions. The outcome of this reaction depends on the coordination mode of the versatile ligand L . Interaction of CO2 with the isomer 1‐π , in which L binds to Fe through the phosphinine moiety in an η5 fashion, leads to the formation of 3‐π , in which CO2 has undergone electrophilic addition to the phosphinine group. In contrast, interaction with 1‐σ —in which L acts as a σ‐chelating [P,N] ligand—leads to product 3‐σ in which one C=O bond has been completely broken. Such CO2 cleavage reactions are extremely rare for late 3d metals, and this represents the first such example mediated by a single Fe centre.  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