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.
2.
p-Fluorophenylisocyanide (CNPhpF) reacts with [Re(CO)5Br] under stepwise exchange of the carbonyl ligands depending on the conditions applied. The reaction stops with the formation of fac-[Re(CO)3Br(CNPhpF)2] in boiling THF. An ongoing carbonyl exchange is observed at higher temperatures, e. g. in refluxing toluene, with the final formation of the [Re(CNPhpF)6]+ cation. The progress of the reactions has been studied by 19F NMR spectroscopy and the structures of [Re(CO)Br(CNPhpF)4] and [Re(CNPhpF)6](BPh4) have been elucidated by X-ray diffraction.  相似文献   

3.
Neutral oxorhenium(V) complexes with thiosemicarbazones derived from 2‐pyridine formamide, HL1, are formed when [ReOCl3(PPh3)2] reacts with equimolar amounts of the ligands. Reduction of the metal and the formation of rhenium(III) complexes of the composition [Re(L1)2]+ occurs when an excess of thiosemicarbazones is used and the reaction is performed in boiling toluene for a prolonged period of time. The thiosemicarbazones deprotonate and act as tridentate ligands as has been confirmed by an X‐ray structure of [ReOCl2(L1b)], where HL1b is 2‐pyridineformamide‐N(4)‐ethylthiosemicarbazone and the ligand occupies the equatorial coordination sphere of the complex together with one of the chloro ligands.  相似文献   

4.

The reaction of a two-fold molar excess of the potential N,O-donor ligand 2-(hydroxymethyl)-1-methylimidazole (Hmi) with trans-[ReOCl3(PPh3)2] led to the isolation of cis-[ReOCl2(mi)(PPh3)]. An X-ray structure determination indicated that the complex has distorted octahedral geometry, and that mi coordinates as a bidentate with the alcoholate oxygen trans to the oxo group. A similar reaction with 2-(1-ethyloxomethyl)-1-methylimidazole (eomi), the ethyl substituted analogue of Hmi, led to the formation of the oxo-bridged dinuclear complex [(μ-O){ReOCl2(eomi)2}2]. The ligand eomi coordinates as a monodentate via the imidazole nitrogen, with the "hard" ether oxygen uncoordinated. An X-ray crystal structure indicates that the chlorides are trans to each other in the ReN2Cl2 planes, which are orthogonal to the O=Re-O-Re=O backbone.  相似文献   

5.
[Tc(NPh)Cl3(PPh3)2] or [Re(NPh)Cl3(PPh3)2] react with two equivalents of Na2mnt (mnt2– = 1,2‐dicyanoethene‐1,2‐dithiolate) with formation of anionic complexes of the composition [M(NPh)(mnt)2]. The products can be isolated as large red blocks of their AsPh4+ salts. The complex anions contain square‐pyramidal coordinated metal atoms with the phenylimido ligands in apical positions. The M–N–C bonds are almost linear. A similar phenylimido complex with an additional amino group was synthesized from [Re(NC6H4‐4‐NH2)Cl3(PPh3)2]. The presence of such substituents may allow coupling of the metal complexes to biomolecules such as peptides, proteins, or sugars, provided the M=N bonds are sufficiently stable against hydrolysis.  相似文献   

6.
The synthesis and characterization of the first two Re complexes with semicarbazone ligands is presented. Selected ligands are 5‐Nitro‐2‐furaldehyde semicarbazone (Nitrofurazone) ( L1 ) and its derivative 3‐(5‐Nitrofuryl)acroleine semicarbazone ( L2 ). Complexes of general formula [ReVOCl2(PPh3) L ], where L = L1 and L2 , were prepared in good yields and high purity by reaction of [ReVOCl3(PPh3)2] with L in ethanol or methanol solutions. The complexes formula and molecular structures were supported by elemental analyses and electronic, FTIR, 1H, 13C and 31P NMR spectroscopies. In addition, the crystal and molecular structure of [ReVOCl2(PPh3) L2 ] was determined by X‐ray diffraction methods. [ReOCl2(PPh3)(3‐(5‐Nitrofuryl)acroleine semicarbazone)] crystallizes in the space group P‐1 with a = 11.2334(2), b = 11.3040(2), c = 12.5040(2) Å, α = 81.861(1), β = 63.555(1), γ = 83.626(1)°, and Z = 2. The Re(V) ion is in a distorted octahedral environment, equatorially coordinated to a deprotonated semicarbazone molecule acting as a bidentate ligand through its carbonylic oxygen and azomethynic nitrogen atoms, to an oxo ligand and a chlorine atom. The six‐fold coordination is completed by another chlorine atom and a triphenylphosphine ligand at the axial positions.  相似文献   

7.
[ReOCl3(PPh3)2] and [Re(NPh)Br3(PPh3)2] react at room temperature with equivalent amounts of N,N‐dialkyl‐N′‐benzoylthioureas (HR1R2btu) in CH2Cl2 under formation of the rhenium(V) complexes [ReOCl2(R1R1btu)(PPh3)] and [Re(NPh)Br2(R1R2btu)(PPh3)], respectively. The products are structurally analogous with the oxygen atoms of the benzoylthioureas binding in trans positions to the oxo or phenylimido ligands. Prolonged reaction times result in the reduction of the oxo compound by the released PPh3 and the formation of rhenium(III) complexes of the composition [ReCl2(PPh3)2(R1R2btu)], while such a second reaction path is excluded for the phenylimido compound. Phenylimido species with more than one N,N‐dialkyl‐N′‐benzoylthioureato ligand could not be isolated, even when a large excess of HR1R2btu was used during the reaction.  相似文献   

8.
Ketenimine complexes are readily available in great variety by reaction of isocyanides with carbene complexes. They have proven to be useful building blocks in new synthetic approaches to carbocyclic and N-heterocyclic four-, five-, and six-membered rings. The reactions involve new metal-induced bond formation patterns of the ketenimine ligands, which can be influenced across a wide range by varying the following five parameters: the metal, the ligands, and the three substituents on the N?C?C unit.  相似文献   

9.
10.
Dimeric, neutral rhenium(I) complexes of the composition [Re2(CO)6X(LR)] (X = Cl or Br) are formed when [NEt4]2[Re(CO)3Br3] or [Re(CO)3Cl(CH3CN)2] react with 2, 2′-dipyridylketone thiosemicarbazones (HLR, R = H, Ph). The thiosemicarbazones deprotonate during the reaction and connect two tricarbonylrhenium(I) units as formally pentadentate bridging ligands. This results in an extremely rare coordination mode, in which the two nitrogen atoms of the hydrazone unit bind to each one of the rhenium atoms. The bond lengths inside the thiosemicarbazonato backbone reflect a large degree of delocalization of electron density.  相似文献   

11.
《Analytical letters》2012,45(15):2329-2342
The synthesis, characterization, and mass spectra of oxorhenium(V) complexes with 1,2-dihydroxybenzene, 1,2,3-trihydroxybenzene, and 2,3-dihydroxynaphtalene are reported. Electrospray ionization, atmospheric pressure photoionization, and laser desorption/ionization mass spectra of the complexes showed abundant negatively charged molecular anions and low fragmentation. Calculated similarity indexes showed significant conformity between the computed and experimental isotopic patterns of selected ions and confirmed correct assignment of elemental composition to m/z values. Electrospray tandem mass spectrometry provided essential information about fragments from molecular ions of studied complexes, making it possible to distinguish among fragment ions and the ions arising from compounds present in the reaction mixture. Based on the results, mass spectrometry utilizing soft common ionization techniques is useful for monitoring complex formation reaction kinetics and the stabilities of the complexes. Representative spectra were recorded for micromolar concentrations of the analytes.  相似文献   

12.
[NEt4]2[Re(CO)3Br3] and [NEt4]2[Tc(CO)3Cl3] react with trimethylsilyltriphenylphosphoraneimine, Me3SiNPPh3, under exchange of the bromo ligands and the formation of cationic [M(CO)3(HNPPh3)3]+ complexes (M = Re, Tc). The required protons are abstracted from the solvent CH2Cl2. The steric bulk of the organic ligands causes a marked distortion of the established coordination polyhedra from an idealized octahedron with bond angles between neighbouring donor atoms between 81.81(8)° and 96.66(8)°. The reaction of [NEt4]2[Re(CO)3Br3] with Me3SiNP(Ph2)CH2PPh2 in CH2Cl2 yields the neutral complex [Re(CO)3Br{HNP(Ph2)CH2PPh2)], which contains a neutral, chelate‐bonded (diphenylphosphinomethyl)diphenylphosphoraneimine ligand. A similar reaction with the bifunctional phosphoraneimine Me3SiNP(Ph2)CH2(Ph2)PNSiMe3 gives only small amounts of a binuclear rhenium(I) complex of the composition [{Re(CO)3Br2}2(HNP(Ph2)CH2(Ph2)PNH)]2‐, whereas the major amount of the bis‐phosphoraneimine undergoes an intramolecular rearrangement to yield [H2NP(Ph2)NP(Ph2)CH3]Br. An X‐ray structure analysis shows a widespread delocalization of electron density over the central part of the cation.  相似文献   

13.
Reactions of 2-(diphenylphosphinomethyl)aniline, H2L1, with [MNCl2(PPh3)2] complexes (M = Re, Tc) give the bis-chelates [MNCl(H2L1)2]Cl (M = Re, Tc) or the mono-chelate [ReNCl2(PPh3)(H2L1)] depending on the conditions applied. The aminophosphine reacts as a bidentate, neutral ligand in all three cases. The complexes were studied spectroscopically and by X-ray crystallography.  相似文献   

14.
Hydrotris(3, 5‐dimethylpyrazol‐1‐yl)borate and hydrotris(3‐phenylpyrazol‐1‐yl)borate decompose during reactions with [ReOCl3(PPh3)2] and [NEt4]2[Re(CO)3Br3], respectively. The generated pyrazole ligands form complexes with the rhenium(V) oxo and the rhenium(I ) tricarbonyl cores. X‐ray crystal structures of the oxo‐bridged dimer [Cl(PPh3)(O)Re(μ‐O)(μ‐Me2pz)2Re(O)(HMe2pz)Cl] ( 1 ) and [Re(CO)3(HPhpz)2(Phpz)] ( 2 ) (HMe2pz = 3, 5‐dimethylpyrazole, HPhpz = 3‐phenylpyrazole) show that the substituted pyrazoles can readily deprotonate and act as monodentate or bridging anionic ligands. Re‐N bond lengths between 2.09 and 2.14Å have been observed for the bridging and between 2.12 and 2.23Å for the terminal pyrazole ligands.  相似文献   

15.
Novel rhenium(I) tricarbonyl complexes have been prepared by reactions of (Et4N)2[Re(CO)3Br3] with acetylpyridine benzoylhydrazone, Hapbhyd, di(2‐pyridyl)ketone benzoylhydrazone, Hpy2bhyd, bis(2‐pyridine)ketone, py2CO, and pyridinealdehyde terephtalaldehydebishydrazone, pytehyd. The ligands remain protonated when no supporting base is added and the following complexes have been isolated: [Re(CO)3Br(Hapbhyd)], [Re(CO)3Br(Hpy2bhyd‐py, hyd)], [Re(CO)3Br(Hpy2bhyd‐py1, py2)], [Re(CO)3Br(py2CO‐N, N)] and [Re(CO)3Br(pytehyd)]. Addition of triethyl amine results in deprotonation of Hapbhyd and the formation of [Re(CO)3(OH2)(apbhyd)], whereas Hpy2bhyd is hydrolysed and a rhenium complex with the monoanionic bis(2‐pyridyl)hydroxymethanolato ligand, {py2C(OH)O}, is formed. The same compound, [Re(CO)3{py2C(OH)O}], is obtained when triethyl amine and water are added to a mixture of (Et4N)2[Re(CO)3Br3] and py2CO. The air‐stable products have been studied by spectroscopic methods and X‐ray crystallography.  相似文献   

16.
Starting from their six-coordinate iron(II) precursor complexes [L8RFe(MeCN)]2+, a series of iron(III) complexes of the known macrocyclic tetracarbene ligand L8H and its new octamethylated derivative L8Me, both providing four imidazol-2-yliden donors, were synthesized. Several five- and six-coordinate iron(III) complexes with different axial ligands (Cl, OTf, MeCN) were structurally characterized by X-ray diffraction and analyzed in detail with respect to their spin state variations, using a bouquet of spectroscopic methods (NMR, UV/Vis, EPR, and 57Fe Mößbauer). Depending on the axial ligands, either low-spin (S=1/2) or intermediate-spin (S=3/2) states were observed, whereas high-spin (S=5/2) states were inaccessible because of the extremely strong in-plane σ-donor character of the macrocyclic tetracarbene ligands. These findings are reminiscent of the spin state patterns of topologically related ferric porphyrin complexes. The ring conformations and dynamics of the macrocyclic tetracarbene ligands in their iron(II), iron(III) and μ-oxo diiron(III) complexes were also studied.  相似文献   

17.
The synthesis of new tripodal nitrogen ligands derived from tris(pyrazolyl)methane (TpmR, R = H, tBu, Ph in 3‐position) is described. After deprotonation of the parent tris(pyrazolyl)methane TpmR, the carbanion reacts readily with ethylene oxide to yield the 3,3,3‐tris(3′‐substituted pyrazolyl)propanol ligands[(3‐Rpz)3CCH2CH2OH, R = H, tBu, Ph, 1a – c ]. These ligands can be easily derivatised at the alcohol function. Microwave‐assisted reactions of these ligands and [Re(CO)5Br] yields the complex [( 1a )Re(CO)3]Br ( 4 ) in the case of ligand 1a , whereas in the case of the substituted ligands 1b and 1c degradation was observed. The degradation products are identified as [(HpzR)2Re(CO)3Br] [R = tBu ( 7b ), Ph ( 7c )]. These complexes were also prepared directly from [Re(CO)5Br] and the corresponding pyrazoles by microwave‐assisted synthesis. The Re(CO)3 complexes 4 and [( 1a )Re(CO)3]OTf ( 5 ) are water‐soluble. The structures of 5· H2O and [{(pz)3CCH2CH3}Re(CO)3]OTf · 1.5H2O · 1/2CH3CN ( 6· 1.5H2O · 1/2CH3CN) as well as the structure of 7b have been elucidated by X‐ray crystallography.  相似文献   

18.
The reaction of [ReBr(CO)5] with phosphite and phosphonite ligands in toluene yielded cis, mer‐[ReBr(CO)2L3] ( 2 : L = P(OMe)3 2a : P(OEt)3 2b : PPh(OMe)2 2c : PPh(OEt)2 2d ). Compounds 2c and 2d were also obtained, as were the phosphinite complexes 2e [L = PPh2(OMe)] and 2f [L = PPh2(OEt)], by reaction of the corresponding phosphorus ligand with trans, mer‐[ReBr(CO)3L2]. Compounds 2 were all characterized by elemental analysis, mass spectrometry and NMR spectroscopy, and the structures of 2b , 2c and 2d were determined by X‐ray diffractometry. Compounds 2a‐d are stable in chloroform and dichloromethane, but 2e and 2f are transformed into the corresponding trans, mer‐[ReBr(CO)3L2] complexes by a reaction for which a partial mechanism is put forward.  相似文献   

19.
The synthesis and characterisation of a homologous series of rhodium 2,2′-biphenyl complexes featuring intramolecular dative bonding of the nominally inert and weakly coordinating trifluoromethyl group are described. Presence of these interactions is evidenced in the solid state using X-ray diffraction, with Rh−F contacts of 2.36–2.45 Å, and in solution using NMR spectroscopy, through hindered C−CF3 bond rotation and the presence of time-averaged 1JRhF and 2JPF coupling.  相似文献   

20.
A series of tricarbonyl rhenium(I) complexes of the type fac‐[ReI(CO)3(ppl)(L)]0/+, where ppl is pyrazino[2,3‐f][1,10]phenanthroline, and where L is Cl?, TfO?, 4‐(tert‐butyl)pyridine (tBu‐py), 4‐methoxypyridine (MeO‐py), 4,4′‐bipyridyl (bpy), or 10‐(picolin‐4‐yl)phenothiazine (pptz), were synthesized and fully characterized. In all complexes, an increment in the electron‐acceptor properties of ppl compared to the free ligand was observed. This effect was more significant for pyridine‐type ligands, especially for pptz, compared to Cl? or TfO?. The properties of fac‐[Re(CO)3(ppl)(pptz)]PF6 were compared with those of the analogous compound fac‐[Re(CO)3(dppz)(pptz)]PF6, where dppz is dipyrido(3,2‐a : 2′,3′‐c)phenazine, the goal being to generate long‐lived excited charge‐transfer (CT) states. In this respect, fac‐[Re(CO)3(ppl)(pptz)]PF6 seems to be a promising candidate.  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