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 15 毫秒
1.
N-heterocyclic bis-carbene ligand (bis-NHC) which was derived from 1,1′-diisopropyl-3,3′-ethylenediimidazolium dibromide (L·2HBr) via silver carbene transfer method, reacted with [(η6-p-cymene)RuCl2]2 and [CpMCl2]2 (Cp = η5-C5Me5, M = Ir, Rh) respectively, afforded complexes [(η6-p-cymene)RuCl2]2(L) (1), [CpIrCl2]2(L) (2) and [CpRhCl(L)][CpRhCl3] (3). When [CpIrCl2]2 was treated with 2 equiv AgOTf at first, and then reacted with bis-NHC ligand, [CpIrCl(L)]OTf (4) was obtained. The molecular structures of complexes 1-4 were determined by X-ray single crystal analysis, showing that 1 and 2 adopted bridging coordination mode, 3 and 4 adopted chelating coordination mode. All of these complexes were characterized by 1H, 13C NMR spectroscopy and element analysis.  相似文献   

2.
Reactions of the dinuclear complexes [(η6-arene)Ru(μ-Cl)Cl]2 (arene = C6H6, p-iPrC6H4Me) and [(η5-C5Me5)M(μ-Cl)Cl]2 (M = Rh, Ir) with 2-substituted-1,8-naphthyridine ligands, 2-(2-pyridyl)-1,8-naphthyridine (pyNp), 2-(2-thiazolyl)-1,8-naphthyridine (tzNp) and 2-(2-furyl)-1,8-naphthyridine (fuNp), lead to the formation of the mononuclear cationic complexes [(η6-C6H6)Ru(L)Cl]+ {L = pyNp (1); tzNp (2); fuNp (3)}, [(η6-p-iPrC6H4Me)Ru(L)Cl]+ {L = pyNp (4); tzNp (5); fuNp (6)}, [(η5-C5Me5)Rh(L)Cl]+ {L = pyNp (7); tzNp (8); fuNp (9)} and [(η5-C5Me5)Ir(L)Cl]+ {L = pyNp (10); tzNp (11); fuNp (12)}. All these complexes are isolated as chloro or hexafluorophosphate salts and characterized by IR, NMR, mass spectrometry and UV/Vis spectroscopy. The molecular structures of [1]Cl, [2]PF6, [4]PF6, [5]PF6 and [10]PF6 have been established by single crystal X-ray structure analysis.  相似文献   

3.
The reactions of 1 mol equiv. each of [Ru(PPh3)3Cl2] and N-(acetyl)-N′-(5-R-salicylidene)hydrazines (H2ahsR, R = H, OCH3, Cl, Br and NO2) in alcoholic media afford simultaneously two types of complexes having the general formulae [Ru(HahsR)(PPh3)2Cl2] and [Ru(ahsR)(PPh3)2Cl]. The complexes have been characterized by elemental analysis, magnetic, spectroscopic and electrochemical measurements. Molecular structures of [Ru(HahsH)(PPh3)2Cl2] and [Ru(ahsH)(PPh3)2Cl] have been confirmed by X-ray crystallography. In both species, the PPh3 ligands are trans to each other. The bidentate HahsH coordinates to the metal ion via the O atom of the deprotonated amide and the imine–N atom in [Ru(HahsH)(PPh3)2Cl2]. In HahsH, the phenolic OH is involved in a strong intramolecular hydrogen bond with the uncoordinated amide N atom forming a seven-membered ring. In [Ru(ahsH)(PPh3)2Cl], the tridentate ahsH2− binds to the metal ion via the deprotonated amide O, the imine N and the phenolate O atoms. In the electronic spectra, the green [Ru(HahsR)(PPh3)2Cl2] and brown [Ru(ahsR)(PPh3)2Cl] complexes display several absorptions in the ranges 385–283 and 457–269 nm, respectively. Both complexes are low-spin and display rhombic EPR spectra in frozen solutions. Both types of complexes are redox active and display a quasi-reversible ruthenium(III) to ruthenium(II) reduction which is sensitive to the polar effect of the substituent on the chelating ligand. The reduction potentials are in the ranges −0.21 to −0.12 and −0.42 to −0.21 V (versus Ag/AgCl) for [Ru(HahsR)(PPh3)2Cl2] and [Ru(ahsR)(PPh3)2Cl], respectively.  相似文献   

4.
Two binuclear complexes [CpM(Cl)CarbS]2 (Cp = η5-C5Me5, M = Rh (1a), CarbS = SC2(H)B10H10, Ir (1b)) were synthesized by the reaction of LiCarbS with the dimeric metal complexes [CpMCl(μ-Cl)]2 (M = Rh, Ir). Four mononuclear complexes CpM(Cl)(L)CarbS (L = BunPPh2, M = Rh (2a), Ir (2b); L = PPh3, M = Rh (4a), Ir (4b)) were synthesized by reactions of 1a or 1b with L (L = BunPPh2 (2); PPh3 (4)) in moderate yields, respectively. Complexes 3a, 3b, 5a, 5b were obtained by treatment of 2a, 2b, 4a, 4b with AgPF6 in high yields, respectively. All of these compounds were fully characterized by IR, NMR, and elemental analysis, and the crystal structures of 1a, 1b, 2a, 2b, 4a, 4b were also confirmed by X-ray crystallography. Their structures showed 3a, 3b and 5a, 5b could be expected as good candidates for heterolytic dihydrogen activation. Preliminary experiments on the dihydrogen activation driven by these half-sandwich Rh, Ir complexes were done under mild conditions.  相似文献   

5.
The syntheses and crystal structures of four new uranyl complexes with [O,N,O,N′]-type ligands are described. The reaction between uranyl nitrate hexahydrate and the phenolic ligand [(N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-N′,N′-dimethylethylenediamine)], H2L1 in a 1:2 molar ratio (M to L), yields a uranyl complex with the formula [UO2(HL1)(NO3)] · CH3CN (1). In the presence of a base (triethylamine, one mole per ligand mole) with the same molar ratio, the uranyl complex [UO2(HL1)2] (2) is formed. The reaction between uranyl nitrate hexahydrate and the ligand [(N,N-bis(2-hydroxy-3,5-di-t-butylbenzyl)-N′,N′-dimethylethylenediamine)], H2L2, yields a uranyl complex with the formula [UO2(HL2)(NO3)] · 2CH3CN (3) and the ligand [N-(2-pyridylmethyl)-N,N-bis(2-hydroxy-3,5-dimethylbenzyl)amine], H2L3, in the presence of a base yields a uranyl complex with the formula [UO2(HL3)2] · 2CH3CN (4). The molecular structures of 14 were verified by X-ray crystallography. The complexes 14 are zwitter ions with a neutral net charge. Compounds 1 and 3 are rare neutral mononuclear [UO2(HLn)(NO3)] complexes with the nitrate bonded in η2-fashion to the uranyl ion. Furthermore, the ability of the ligands H2L1–H2L4 to extract the uranyl ion from water to dichloromethane, and the selectivity of extraction with ligands H2L1, H3L5 (N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-3-amino-1-propanol), H2L6 · HCl (N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-1-aminobutane · HCl) and H3L7 · HCl (N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-6-amino-1-hexanol · HCl) under varied chemical conditions were studied. As a result, the most efficient and selective ligand for uranyl ion extraction proved to be H3L7 · HCl.  相似文献   

6.
7.
Reactions of [{M(μ‐Cl)(coe)2}2] (M = Rh, Ir; coe = cis‐cyclooctene) with the secondary phosphane tBu2PH under various molar ratios were investigated. Probably, for kinetic reasons, the reaction behavior of the rhodium species differed from that of the iridium analogue in some instances. During these studies complexes [MCl(tBu2PH)3] [M = Rh ( 1 ), Ir ( 2 )] were isolated, and solution variable‐temperature 31P{1H} NMR studies revealed that these complexes show a conformational rigidity on the NMR time scale. Spectra recorded in the temperature range from 173 to 373 K indicated in each case only one rotamer containing three chemically nonequivalent phosphanes due to the restricted rotation of these ligands about the M–P bonds and the tert‐butyl substituents around the P–C(tBu) bonds, respectively. Compound 1 showed in solution already at room temperature in several solvents a dissociation of a phosphane ligand affording the known complex [{Rh(μ‐Cl)(tBu2PH)2}2] beside the free phosphane. In contrast to these findings, the iridium analogue 2 remained completely unchanged under similar conditions and exhibited, therefore, some kinetic inertness. For a better understanding of the NMR spectroscopic investigations, the molecular structure of 1 in the solid state was confirmed by X‐ray crystallography.  相似文献   

8.
Reactions of Li2[PdCl4] with N-(aroyl)-N′-(2,4-dimethoxybenzylidene)hydrazines (H2L = (4-R-C6H4)C(O)NHNCH(2,4-(CH3O)2C6H3)) in presence of PPh3 have provided cyclopalladated complexes having the general formula [PdL(PPh3)] (1, 2 and 3 where R = OCH3, CH3 and Cl, respectively) and a coordination complex trans-[Pd(HL)(PPh3)2Cl] (4 where R = NO2). The complexes have been characterized by elemental analysis, infrared, 1H NMR and electronic absorption spectroscopy. X-ray structures of all the complexes have been determined. In 1, 2 and 3, the C,N,O-donor dianionic ligand (L2−) forms two fused five-membered chelate rings at the metal centre. On the other hand, the monoanionic ligand (HL) acts as deprotonated amide N-donor in 4. The strong electron withdrawing effect of the nitro group on the aroyl fragment is possibly responsible for the monodentate amide N-coordinating behavior of HL in 4. The orientation of the 4-methoxy group in 1 is different than that in 2 and 3 due to intramolecular C-H?π interaction in the last two complexes. The structure of 4 shows an apical C-H?Pd interaction involving the azomethine (-CHN-) group of HL. In the crystal lattice of all the structures, various types of intermolecular non-covalent interactions are present. The self-assembly of the molecules of 1, 2 and 3 leads to two-dimensional networks. The same network observed for 2 and 3 reflects the interchangeability of the chloro and methyl groups due to their similar volumes. In the case of 4, the complex and the water molecules present in the crystal lattice form parallel homo-chiral helices and finally interhelical interactions lead to a three-dimensional network.  相似文献   

9.
The syntheses, structures, spectroscopy, and electrochemistry for six Ir(III) and Rh(III) mixed sandwich mononuclear complexes involving tridentate macrocycles and pentamethylcyclopentadienide (Cp*) are reported. The complexes are readily prepared by direct ligand substitution reactions from the dichloro bridged binuclear complexes, [{M(Cp*)(Cl)2}2]. All complexes have the general formula [M(L)(Cp*)]X2 (M = Ir(III) or Rh(III), L = macrocycle, or Cl) and exhibit a distorted octahedral structure involving three donor atoms from the macrocycle and the facially coordinating carbocyclic Cp* ligand. The complex cations include: [Rh(η5 -Cp*)(9S3)]2+ (1), [Rh(η5-Cp*)(9N3)]2+ (2), [Rh(η5-Cp*)(10S3)]2+ (3), [Ir(η5-Cp*)(9S3)]2+ (4), [Ir(η5-Cp*)(9N3)]2+ (5), and [Ir(η5-Cp*)(10S3)]2+ (6), where 9S3 = 1,4,7-trithiacyclononane, 9N3 = 1,4,7-triazacyclononane, and 10S3 = 1,4,7-trithiacyclodecane. The structures for all six complexes are supported by 1H and 13C{1H} NMR spectroscopy, and five complexes are also characterized by single-crystal X-ray crystallography (complexes 1-5). The 1H NMR splittings between the two sets of methylene protons for both the Rh(III) and Ir(III) 9S3 complexes are much larger (0.4 vs. 0.2 ppm) compared to those in the two 9N3 complexes. Similarly, the 13C{1H} NMR spectra in all four thioether complexes show that the ring carbons in the Cp* ligand are shifted by over 10 ppm downfield compared to the azacrown complexes. The electrochemistry of the complexes is surprisingly invariable and is dominated by a single irreversible metal-centered reduction near −1.2 V vs. Fc/Fc+.  相似文献   

10.
Half-sandwich complexes of formula [(ηn-ring)MClL]PF6 [L = (S)-2-[(Sp)-2-(diphenylphosphino)ferrocenyl]-4-isopropyloxazoline; (ηn-ring)M = (η5-C5Me5)Rh; (η5-C5Me5)Ir; (η6-p-MeC6H4iPr)Ru; (η6-p-MeC6H4iPr)Os] have been prepared and spectroscopically characterised. The molecular structures of the rhodium and iridium compounds have been determined by X-ray crystallography. The related solvate complexes [(η5-C5Me5)ML(Me2CO)]2+ (M = Rh, Ir) are active catalysts for the Diels-Alder reaction between methacrolein and cyclopentadiene.  相似文献   

11.
The synthesis and characterization of heteroleptic complexes with the formulations [(η6-arene)RuCl(fcdpm)] (η6-arene = C6H6, C10H14) and [(η5-C5Me5)MCl(fcdpm)] (M = Rh, Ir; fcdpm = 5-ferrocenyldipyrromethene) have been reported. All the complexes have been characterized by elemental analyses, IR, 1H NMR and electronic spectral studies. Structures of [(η6-C6H6)RuCl(fcdpm)] and [(η6-C10H14)RuCl(fcdpm)] have been determined crystallographically. Chelating monoanionic linkage of fcdpm to the respective metal centres has been supported by spectral and structural studies. Further, reactivity of the representative complex [(η6-C10H14)RuCl(fcdpm)] with ammonium thiocyanate (NH4SCN) and triphenylphosphine (PPh3) have been examined.  相似文献   

12.
The syntheses and structural studies of an [O,N,O,N′]-type phenolic ligand [(N’,N’-bis(2-hydroxy-3-methoxy-5-(propen-2-yl)benzyl)-N-(2-aminoethyl)morpholine), (H2L) and two new uranyl complexes of this ligand are described. The reaction between uranyl nitrate hexahydrate and H2L in a 1:2 M ratio (M to H2L) results in a uranyl complex of the formula [UO2(HL)(NO3)(H2O)] (1). In the presence of a base (triethylamine), with the same molar ratio, the uranyl complex [UO2(HL)2]·2CH3CN (2) is formed. The molecular structures H2L, 1 and 2 were verified by X-ray crystallography. Both uranyl complexes are zwitterions with a neutral net charge. A comprehensive NMR-structural analyses of all compounds were performed in CDCl3, DMSO-d6 and pyridine-d5. Complex 2 dissociates in all the studied NMR-solvents, forming a 1:1 complex and free ligand, but according to the spectra the formed complexes are not alike. The results of the ability of the ligand to extract the uranyl ion from water into dichloromethane are also presented.  相似文献   

13.
A study has been undertaken of the lithium and sodium metallation of N,N′-di(aryl)formamidines with alkyl groups at the 2- and 6-aryl position. 1H NMR and X-ray diffraction data indicate an increase in steric bulk from 2,6-dimethylphenyl to 2,6-diethylphenyl to 2,6-diisopropylphenyl incites incremental changes in amidinate binding and nuclearity. In selected instances, this invokes the first lithium and sodium amidinate/guanidinate complexes to exhibit metal-arene contacts.  相似文献   

14.
Cyclopalladated complexes with the Schiff base N-(benzoyl)-N-(2,4-dimethoxybenzylidene)hydrazine (H2L, 1) have been described. The reaction of 1 with Li2[PdCl4] in methanol yields the complex [Pd(HL)Cl] (2). [Pd(HL)(CH3CN)Cl] (3) has been prepared by dissolving 2 in acetonitrile. In methanol-acetonitrile mixture, treatment of 2 with two mole equivalents of PPh3 produces [PdL(PPh3)] (4) and that with one mole equivalent of PPh3 produces [Pd(HL)(PPh3)Cl] (5). Crystallization of 2 from dmso-d6 results into isolation of [Pd(HL)((CD3)2SO)Cl] (6). In 2, the monoanionic ligand (HL) is C,N,O-donor and the Cl-atom is trans to the azomethine N-atom. In 3, 5 and 6, HL is C,N-donor and the Cl-atom is trans to the metallated C-atom. The remaining fourth coordination site is occupied by the N-atom of CH3CN, the P-atom of PPh3 and the S-atom of (CD3)2SO in 3, 5 and 6, respectively. Thus on dissolution in acetonitrile and dmso and in reaction with stoichiometric PPh3 the incoming ligand imposes a rearrangement of the coordinating atoms on the palladium centre. On the other hand, in presence of excess PPh3 deprotonation of the amide functionality in 2 occurs and the Cl-atom is replaced by the P-atom of PPh3 to form 4. Here the dianionic ligand (L2−) remains C,N,O-donor as in 2. The compounds have been characterized with the help of elemental analysis (C, H, N), infrared, 1H NMR and electronic absorption spectroscopy. Molecular structures of 3, 4, and 6 have been determined by X-ray crystallography.  相似文献   

15.
N-2-(3-picolyl)-N′-phenylthiourea, 3PicTuPh, monoclinic, P21/n, a=7.617(2) b=7.197(5), c=22.889(5) Å, β=94.63(4)°, V=1250.7(1) Å3 and Z=4; N-2-(4-picolyl)-N′-phenylthiourea, 4PicTuPh, triclinic, P-1, a=7.3960(5), b=7.9660(12), c=21.600(3) Å, α=86.401(4), β=84.899(8), γ=77.769(8)°, V=1237.5(3) Å3 and Z=4; N-2-(5-picolyl)-N′-phenylthiourea, 5PicTuPh, monoclinic, P21/c, a=14.201(1), b=4.905(3), c=17.689(3) Å, β=91.38(1)°, V=1231.8(7) Å3 and Z=4; N-2-(6-picolyl)-N′-phenylthiourea, 6PicTuPh, monoclinic, C2/c2, a=14.713(1), b=9.367(1), c=18.227(1) Å, β=92.88(1)°, V=2515.5(1) Å3 and Z=8 and N-2-(4,6-lutidyl)-N′-phenylthiourea, 4,6LutTuPh, monoclinic, C2/c, a=11.107(2), b=11.793(2), c=20.084(4) Å, β=96.10(3)°, V=2616(1) Å3 and Z=8. Intramolecular hydrogen bonding between N′H and the pyridyl nitrogen and intermolecular hydrogen bonding involving the thione sulfur are affected by substitution of the pyridine ring, as is the planarity of the molecule. 1H NMR studies in CDCl3 show the NH′ hydrogen resonance considerably downfield from other resonances in the spectrum for each thiourea.  相似文献   

16.
The dinuclear dichloro complexes [(η6-arene)2Ru2(μ-Cl)2Cl2] and [(η5-C5Me5)2M2(μ-Cl)2Cl2] react with 2-(pyridine-2-yl)thiazole (pyTz) to afford the cationic complexes [(η6-arene)Ru(pyTz)Cl]+ (arene = C6H61, p-iPrC6H4Me 2 or C6Me63) and [(η5-C5Me5)M(pyTz)Cl]+ (M = Rh 4 or Ir 5), isolated as the chloride salts. The reaction of 2 and 3 with SnCl2 leads to the dinuclear heterometallic trichlorostannyl derivatives [(η6-p-iPrC6H4Me)Ru(pyTz)(SnCl3)]+ (6) and [(η6-C6Me6)Ru(pyTz)(SnCl3)]+ (7), respectively, also isolated as the chloride salts. The molecular structures of 4, 5 and 7 have been established by single-crystal X-ray structure analyses of the corresponding hexafluorophosphate salts. The in vitro anticancer activities of the metal complexes on human ovarian cancer cell lines A2780 and A2780cisR (cisplatin-resistant), as well as their interactions with plasmid DNA and the model protein ubiquitin, have been investigated.  相似文献   

17.
Reactions of [Cp*M(μ-Cl)Cl]2 (M = Ir, Rh; Cp* = η5-pentamethylcyclopentadienyl) with bi- or tri-dentate organochalcogen ligands Mbit (L1), Mbpit (L2), Mbbit (L3) and [TmMe] (L4) (Mbit = 1,1′-methylenebis(3-methyl-imidazole-2-thione); Mbpit = 1,1′-methylene bis (3-iso-propyl-imidazole-2-thione), Mbbit = 1,1′-methylene bis (3-tert-butyl-imidazole-2-thione)) and [TmMe] (TmMe = tris (2-mercapto-1-methylimidazolyl) borate) result in the formation of the 18-electron half-sandwich complexes [Cp*M(Mbit)Cl]Cl (M = Ir, 1a; M = Rh, 1b), [Cp*M(Mbpit)Cl]Cl (M = Ir, 2a; M = Rh, 2b), [Cp*M(Mbbit)Cl]Cl (M = Ir, 3a; M = Rh, 3b) and [Cp*M(TmMe)]Cl (M = Ir, 4a; M = Rh, 4b), respectively. All complexes have been characterized by elemental analysis, NMR and IR spectra. The molecular structures of 1a, 2b and 4a have been determined by X-ray crystallography.  相似文献   

18.
The meso-pyridyl substituted dipyrromethane ligands 5-(4-pyridyl)dipyrromethane (4-dpmane) and 5-(3-pyridyl)dipyrromethane (3-dpmane) have been employed in the synthesis of a series of complexes with the general formulations [(η6-arene)RuCl2(L)] (η6-arene = C6H6, C10H14) and [(η5-C5Me5)MCl2(L)] (M = Rh, Ir). The reaction products have been characterized by microanalyses and spectral studies and molecular structures of the complexes [(η6-C10H14)RuCl2(4-dpmane)] and [(η5-C5Me5)IrCl2(3-dpmane)] have been determined crystallographically. For comparative studies, geometrical optimization have been performed on the complex [(η5-C5Me5)IrCl2(4-dpmane)] using exchange correlation functional B3LYP. Optimized bond length and angles are in good agreement with the structural data of the complex [(η5-C5Me5)IrCl2(3-dpmane)]. The complexes [(η6-C10H14)RuCl2(3-dpmane)], [(η5-C5Me5)RhCl2(3-dpmane)] and [(η5-C5Me5)IrCl2(3-dpmane)] have been employed as a transfer hydrogenation catalyst in the reduction of aldehydes. It was observed that the rhodium and iridium complexes mentioned above are more effective in this regard in comparison to the ruthenium complex.  相似文献   

19.
The mononuclear cationic complexes [(η6-C6H6)RuCl(L)]+ (1), [(η6-p-iPrC6H4Me)RuCl(L)]+ (2), [(η5-C5H5)Ru(PPh3)(L)]+ (3), [(η5-C5Me5)Ru(PPh3)(L)]+ (4), [(η5-C5Me5)RhCl(L)]+ (5), [(η5-C5Me5)IrCl(L)]+ (6) as well as the dinuclear dicationic complexes [{(η6-C6H6)RuCl}2(L)]2+ (7), [{(η6-p-iPrC6H4Me)RuCl}2(L)]2+ (8), [{(η5-C5H5)Ru(PPh3)}2(L)]2+ (9), [{(η5-C5Me5)Ru(PPh3)}2(L)]2+ (10), [{(η5-C5Me5)RhCl}2(L)]2+ (11) and [{(η5-C5Me5)IrCl}2(L)]2+ (12) have been synthesized from 4,4′-bis(2-pyridyl-4-thiazole) (L) and the corresponding complexes [(η6-C6H6)Ru(μ-Cl)Cl]2, [(η6-p-iPrC6H4Me)Ru(μ-Cl)Cl]2, [(η5-C5H5)Ru(PPh3)2Cl)], [(η5-C5Me5)Ru(PPh3)2Cl], [(η5-C5Me5)Rh(μ-Cl)Cl]2 and [(η5-C5Me5)Ir(μ-Cl)Cl]2, respectively. All complexes were isolated as hexafluorophosphate salts and characterized by IR, NMR, mass spectrometry and UV-vis spectroscopy. The X-ray crystal structure analyses of [3]PF6, [5]PF6, [8](PF6)2 and [12](PF6)2 reveal a typical piano-stool geometry around the metal centers with a five-membered metallo-cycle in which 4,4′-bis(2-pyridyl-4-thiazole) acts as a N,N′-chelating ligand.  相似文献   

20.
rac-2-[(Diphenylphosphino)methyl]ferrocenecarboxylic acid (1) was prepared in a good yield from rac-2-(N,N-dimethylaminomethyl)bromoferrocene (2) via rac-2-(hydroxymethyl)bromoferrocene (4) and rac-2-[(diphenylphosphino)methyl]bromoferrocene (5), and further converted to the respective phosphine oxide (6), phosphine sulfide (7) and methyl ester (8). The phosphines 1 and 8 were studied as ligands in rhodium complexes. The reaction of di-μ-chloro-bis[chloro-(η5-pentamethylcyclopentadienyl)rhodium(III)] with the stoichiometric amounts of 1 and 8 yielded the corresponding mononuclear complexes with P-monodentate ligands: [RhC25-C5Me5)(L-κP)], 9 and 10, respectively. Attempted deprotonation of 9 with LiBu or KOt-Bu gave intractable mixtures, in which the parent complex 9 as the major component was accompanied by two new compounds, likely the diastereoizomeric phosphinocarboxylate complexes. A defined O,P-chelating phosphinocarboxylate complex, [SP-4-2]-carbonyl-[rac-2-{(diphenylphosphino)methyl}ferrocenecarboxylato-κ2O,P]-tricyclohexylphosphinerhodium(I) (12), was obtained from the displacement of acetylacetonate(1−) (acac) ligand in [Rh(acac)(CO)(PCy3)] (Cy = cyclohexyl) with acid 1. The structures of 1, 6 · CHCl3, and 7 · 1/2 CH2Cl2, 10, and hydrated complexes 9 and 12 were determined by single-crystal X-ray diffraction.  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