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 62 毫秒
1.
Reaction of π-isopentadienetricarbonyliron (1) with aryl lithium ArLi (Ar?phenyl, p-tolyl, p-methoxyphenyl, p-trifluoromethylphenyl) in ether at low temperature, and subsequent alkylation of the acylmetallate formed with triethyloxonium tetrafluoroborate[(C2H5)3OBF4] in aqueous solution at 0°C, gave orange-red crystalline complexes (2-5), the isomerized products of isopentadiene (dicarbonyl) [ethoxy(aryl)carbene] iron with composition of C5H8 (CO)2FeC(OC2H5)Ar When LiC6Cl5 was used as nucleophilic reagent in the reaction, on alkylation of the afforded acylmetallate intermediate under some reaction conditions, complex (CO)4FeC(OC2H5)C6Cl5 (6) was obtained. The molecular structure of complexes 2 and 6 were determined by means of single crystal X-ray diffraction measurements. IR, 1H NMR and mass spectra of these complexes were investigated.  相似文献   

2.
Imine complexes [IrCl(η5‐C5Me5){κ1‐NH=C(H)Ar}{P(OR)3}]BPh4 ( 1 , 2 ) (Ar = C6H5, 4‐CH3C6H4; R = Me, Et) were prepared by allowing chloro complexes [IrCl25‐C5Me5){P(OR)3}] to react with benzyl azides ArCH2N3. Bis(imine) complexes [Ir(η5‐C5Me5){κ1‐NH=C(H)Ar}2{P(OR)3}](BPh4)2 ( 3 , 4 ) were also prepared by reacting [IrCl25‐C5Me5){P(OR)3}] first with AgOTf and then with benzyl azide. Depending on the experimental conditions, treatment of the dinuclear complex [IrCl25‐C5Me5)]2 with benzyl azide yielded mono‐ [IrCl25‐C5Me5){κ1‐NH=C(H)Ar}] ( 5 ) and bis‐[IrCl(η5‐C5Me5){κ1‐NH=C(H)Ar}2]BPh4 ( 6 ) imine derivatives. In contrast, treatment of chloro complexes [IrCl25‐C5Me5){P(OR)3}] with phenyl azide C6H5N3 gave amine derivatives [IrCl(η5‐C5Me5)(C6H5NH2){P(OR)3}]BPh4 ( 7 , 8 ). The complexes were characterized spectroscopically (IR, NMR) and by X‐ray crystal structure determination of [IrCl(η5‐C5Me5){κ1‐NH=C(H)C6H4‐4‐CH3}{P(OEt)3}]BPh4 ( 2b ).  相似文献   

3.
Crystallization experiments with the dinuclear chelate ring complex di‐μ‐chlorido‐bis[(η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)platinum(II)], [Pt2(C15H19O4)2Cl2], containing a derivative of the natural compound eugenol as ligand, have been performed. Using five different sets of crystallization conditions resulted in four different complexes which can be further used as starting compounds for the synthesis of Pt complexes with promising anticancer activities. In the case of vapour diffusion with the binary chloroform–diethyl ether or methylene chloride–diethyl ether systems, no change of the molecular structure was observed. Using evaporation from acetonitrile (at room temperature), dimethylformamide (DMF, at 313 K) or dimethyl sulfoxide (DMSO, at 313 K), however, resulted in the displacement of a chloride ligand by the solvent, giving, respectively, the mononuclear complexes (acetonitrile‐κN)(η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chloridoplatinum(II) monohydrate, [Pt(C15H19O4)Cl(CH3CN)]·H2O, (η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chlorido(dimethylformamide‐κO)platinum(II), [Pt(C15H19O4)Cl(C2H7NO)], and (η2‐2‐allyl‐4‐methoxy‐5‐{[(propan‐2‐yloxy)carbonyl]methoxy}phenyl‐κC1)chlorido(dimethyl sulfoxide‐κS)platinum(II), determined as the analogue {η2‐2‐allyl‐4‐methoxy‐5‐[(ethoxycarbonyl)methoxy]phenyl‐κC1}chlorido(dimethyl sulfoxide‐κS)platinum(II), [Pt(C14H17O4)Cl(C2H6OS)]. The crystal structures confirm that acetonitrile interacts with the PtII atom via its N atom, while for DMSO, the S atom is the coordinating atom. For the replacement, the longest of the two Pt—Cl bonds is cleaved, leading to a cis position of the solvent ligand with respect to the allyl group. The crystal packing of the complexes is characterized by dimer formation via C—H…O and C—H…π interactions, but no π–π interactions are observed despite the presence of the aromatic ring.  相似文献   

4.
Eight new R1CpTiCl2(OC(C6H4R2)Ph2) complexes were synthesized by the reaction of R1CpTiCl3 with Ph2(R2C6H4)COH (R2C6H4 = phenyl or o‐methyl‐phenyl) in the presence of Et3N in good yield and characterized by 1H NMR, elemental analysis, IR and mass spectrometry. A suitable single crystal of complex 2 (R1: CH3, R2: H) was obtained and the structure determined by X‐ray diffraction. When activated by methylaluminoxane (MAO), all complexes were active for the polymerization of ethylene and styrene. The effect of variation in temperature, catalyst concentration and MAO/catalyst molar ratio was also studied. Complex 5 (R1: n‐C4H9, R2: H) showed a moderate conversion (37.4%) for the polymerization of methyl methacrylate. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

5.
A series of α‐diimine nickel(II) complexes containing chloro‐substituted ligands, [(Ar)N?C(C10H6)C?N(Ar)]NiBr2 ( 4a , Ar = 2,3‐C6H3Cl2; 4b , Ar = 2,4‐C6H3Cl2; 4c , Ar = 2,5‐C6H3Cl2; 4d , Ar = 2,6‐C6H3Cl2; 4e , Ar = 2,4,6‐C6H2Cl3) and [(Ar)N?C(C10H6)C?N(Ar)]2NiBr2 ( 5a , Ar = 2,3‐C6H3Cl2; 5b , Ar = 2,4‐C6H3Cl2; 5c , Ar = 2,5‐C6H3Cl2), have been synthesized and investigated as precatalysts for ethylene polymerization. In the presence of modified methylaluminoxane (MMAO) as a cocatalyst, these complexes are highly effective catalysts for the oligomerization or polymerization of ethylene under mild conditions. The catalyst activity and the properties of the products were strongly affected by the aryl‐substituents of the ligands used. Depending on the catalyst structure, it is possible to obtain the products ranging from linear α‐olefins to high‐molecular weight polyethylenes. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 1964–1974, 2006  相似文献   

6.
Reaction of the novel ruthenacyclopentatriene [(η5-C5H5)Ru(C4Ph2H2)Br] (1) with isocyanides gives the imino-2,5-diphenylcyclopentadiene complexes [(η5-C5H5Ru(η4-C5Ph2H2NR)Br] (2, R = Me, Et, Cy, t-Bu, 2,6-Me2C6H3); a novel fluxional process involving phenyl substituent rotation and imino nitrogen inversion has been identified for 2 (R = t-Bu, 2,6-Me2C6H3), the interpretation of which is supported by the X-ray crystal structure determination of 2 (R = t-Bu).  相似文献   

7.
Treatment of β-diketiminate ligands bearing different N-aryl monoatomic substituents [HLH = (C6H5)N = C(Me)CH=C(Me)NH(C6H5), HLF = (2,6-F2C6H3)N=C(Me)CH=C(Me)NH(2,6-F2C6H3), and HLCl = (2,6-Cl2C6H3)N=C(Me)CH=C(Me)NH(2,6-Cl2C6H3)] with Ln(CH2SiMe3)3(THF)2 (Ln = Y and Lu) afforded a variety of β-diketiminato rare-earth metal complexes depending on substituents, namely, phenyl ring C–H bond activated complexes (L')(LH)Lu(THF) ( 1b , L' = (C6H4)N = C(Me)CH=C(Me)N(C6H5)), six-coordinate homoleptic complexes (LH)3Ln [Ln = Y ( 1aa ), Lu ( 1bb )], five-coordinate monoalkyl complexes (LF)2Ln(CH2SiMe3) [Ln = Y ( 2a ), Lu ( 2b )], and four-coordinate dialkyl complexes (LCl)Ln(CH2SiMe3)2 [Ln = Y ( 3a ), Lu ( 3b )]. All these complexes were characterized with NMR spectroscopy, and lutetium complexes 1b , 1bb and 3b were structurally validated by single-crystal X-ray diffraction analysis. Moreover, dialkyl complexes 3 promoted the polymerization of 2-vinylpyridine (2-VP) to produce atactic poly(2-vinylpyridine) (P2VP) with quantitative yield. On activation with an equimolar amount of [Ph3C][B(C6F5)4], complexes 3 afforded highly isotactic P2VP with an mm value up to 94 %. Both 1H NMR spectrum and MALDI-TOF mass analysis of an oligomer indicate that the polymerization was initiated by coordination insertion of 2-VP into the Y-CH2SiMe3 bond.  相似文献   

8.
A series of new indanimine ligands [ArN?CC2H3(CH3)C6H2(R)OH] (Ar = Ph, R = Me ( 1 ), R = H ( 2 ), and R = Cl ( 3 ); Ar = 2,6‐i‐Pr2C6H3, R = Me ( 4 ), R = H ( 5 ), and R = Cl ( 6 )) were synthesized and characterized. Reaction of indanimines with Ni(OAc)2·4H2O results in the formation of the trinuclear hexa(indaniminato)tri (nickel(II)) complexes Ni3[ArN = CC2H3(CH3)C6H2(R)O]6 (Ar = Ph, R = Me ( 7 ), R = H ( 8 ), and R = Cl ( 9 )) and the mononuclear bis(indaniminato)nickel (II) complexes Ni[ArN?CC2H3(CH3)C6H2(R)O]2 (Ar = 2,6‐i‐Pr2C6H3, R = Me ( 10 ), R = H ( 11 ), and R = Cl ( 12 )). All nickel complexes were characterized by their IR, NMR spectra, and elemental analyses. In addition, X‐ray structure analyses were performed for complexes 7 , 10 , 11 , and 12 . After being activated with methylaluminoxane (MAO), these nickel(II) complexes can polymerize norbornene to produce addition‐type polynorbornene (PNB) with high molecular weight Mv (106 g mol?1), highly catalytic activities up to 2.18 × 107 gPNB mol?1 Ni h?1. Catalytic activities and the molecular weight of PNB have been investigated for various reaction conditions. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 489–500, 2008  相似文献   

9.
The reaction of Au atoms with 12C2H4 or 12C2H4/Ar mixtures at 8–10 K yields a single product. Using Au and 12C2H4 concentration experiments, warm-up studies and 13C2H4/Ar, 12C2H4/13C2H4/Ar isotopic substitution, coupled with infrared and UV-visible spectroscopy, the product is characterized to be monoethylene gold(0), (C2H4)Au, the first reported example of a zerovalent gold-olefin complex. Extended Hückel molecular orbital calculations proved to be a useful aid towards the assignment of the optical spectrum of (C2H4)Au. The thermal stability of (C2H4)Au in solid C2H4 at 70 K is discussed in terms of the feasibility of a macroscale, liquid nitrogen temperature, chemical synthesis. The molecular and electronic properties of the group of complexes (C2H4)M and M(02), where M = Ag or Au, are compared and discussed.  相似文献   

10.
A series of NCO/NCS pincer precursors, 3‐(Ar2OCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCOAr2)Br, 3a , 3b , 3c , 3d ) and 3‐(2,6‐Me2C6H3SCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCSMe)Br, 4a and 4b ) were synthesized and characterized. The reactions of [Ar1NCOAr2]Br/ [Ar1NCSMe]Br with nBuLi and the subsequent addition of the rare‐earth‐metal chlorides afforded their corresponding rare‐earth‐metal–pincer complexes, that is, [(Ar1NCOAr2)YCl2(thf)2] ( 5a , 5b , 5c , 5d ), [(Ar1NCOAr2)LuCl2(thf)2] ( 6a , 6d ), [(Ar1NCOAr2)GdCl2(thf)2] ( 7 ), [{(Ar1NCSMe)Y(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 8 , 9 ), and [{(Ar1NCSMe)Gd(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 10 , 11 ). These diamagnetic complexes were characterized by 1H and 13C NMR spectroscopy and the molecular structures of compounds 5a , 6a , 7 , and 10 were well‐established by X‐ray diffraction analysis. In compounds 5a , 6a , and 7 , all of the metal centers adopted distorted pentagonal bipyramidal geometries with the NCO donors and two oxygen atoms from the coordinated THF molecules in equatorial positions and the two chlorine atoms in apical positions. Complex 10 is a dimer in which the two equal moieties are linked by two chlorine atoms and two Cl? Li? Cl bridges. In each part, the gadolinium atom adopts a distorted pentagonal bipyramidal geometry. Activated with alkylaluminum and borate, the gadolinium and yttrium complexes showed various activities towards the polymerization of isoprene, thereby affording highly cis‐1,4‐selective polyisoprene, whilst the NCO? lutetium complexes were inert under the same conditions.  相似文献   

11.
Electron-rich Phenyl Complexes of Transition Metals. II. Li4Co2(C6H5)4 · 4THF, Li4Co2(C6H5)4 · 3 Dioxan and Li3Co(C6H5)2(LiC6H5) · 5THF, the First Complexes with a Bis(phenyl)-cobalt(0)- and -cobalt(-I) Unity . Li2CoII(C6H5)4 · 4THF reacts spontaneously in benzene by splitting off of two phenyl radicals to a dimeric bis(phenyl) cobalt(0) complex which has been isolated as a THF and a dioxan adduct Li4Co2(C6H5)4 · 4THF and Li4Co2(C6H5)4 · 3 Dioxan, respectively. Reduction with lithiumphenyl in ether gives a phenyl cobalt(-I) complex Li4Co(C6H5)3 · 5THF containing besides σ-bonded phenyl anions lithium phenyl coordinated to cobalt in a π-complex like manner, proved by means of 13C? NMR-spectroscopy. The stabilization of the low oxidation states is explained by coordination of the lithium ions to cobalt by multiple center bonds, and for each compound a plausible structure is derived.  相似文献   

12.
The mono-hydrido-bridged complexes (PEt3)2(Ar)Pt(μ2-H)Pt(Ar)(PEt3)2]-[BPh4] (Ar = Ph, 4-MeC6H4 and 2,4-Me2C6H3) have been obtained by treating trans-[Pt(Ar)(MeOH)(PEt3)2][BF4] with sodium formate and Na[BPH4]. The cations [PEt3)2(Ar)Pt(μ2-H)Pt(Arb')(PEt3)2]b+ (Ar = Ph and Arb' - 2,4-Me2C6H3 and 2,4,6-Me3C6H2 have bee identified in solution. Their b1H- and b31P-NMR data are reported. The X-ray crystal structure of [(PEt3)2(Ph)Pt(μ2-H)Pt(Ph)(PEt3)2][BPh4] is reported.  相似文献   

13.
The known compound phenyltetrafluoroiodine(V) is shown by X-ray diffraction to have a tetragonal pyramidal structure with an apical phenyl group. This structure is compared to that of IF(OTeF5)4, where the apical position is occupied by the fluorine atom. C6H5IF4 adds F, forming C6H5IF5, which has a pentagonal pyramidal structure with an apical phenyl group. Fluoride abstraction from C6H5IF4 by SbF5 results in the formation of the cation C6H5IF3+, which has a pseudotrigonal bipyramidal structure with the phenyl group occupying an equatorial position. Isoelectronic C6H5IOF2 has a similar structure, with the phenyl group and oxygen atom both occupying equatorial positions.  相似文献   

14.
Cationic amidotitanocene complexes [Cp2Ti(NPhAr)][B(C6F5)4] (Cp=η5-C5H5; Ar=phenyl ( 1 a ), p-tolyl ( 1 b ), p-anisyl ( 1 c )) were isolated. The bonding situation was studied by DFT (Density Functional Theory) using EDA-NOCV (Energy Decomposition Analysis with Natural Orbitals for Chemical Valence). The polar Ti−N bond in 1 a–c features an unusual inversion of σ and π bond strengths responsible for the balance between stability and reactivity in these coordinatively unsaturated species. In solution, 1 a–c undergo photolytic Ti−N cleavage to release Ti(III) species and aminyl radicals ⋅ NPhAr. Reaction of 1 b with H3BNHMe2 results in fast homolytic Ti−N cleavage to give [Cp2Ti(H3BNHMe2)][B(C6F5)4] ( 3 ). 1 a–c are highly active precatalysts in olefin hydrogenation and silanes/amines cross-dehydrogenative coupling, whilst 3 efficiently catalyzes amine-borane dehydrogenation. The mechanism of olefin hydrogenation was studied by DFT and the cooperative H2 activation key step was disclosed using the Activation Strain Model (ASM).  相似文献   

15.
The synthesis and structural characterization of two azirine rhodium(III ) complexes are described. The stabilization, N‐coordination and phenylgroup π‐stacking of the highly reactive and strained 3‐phenyl‐2H‐azirine by transition metal coordination is observed. The reaction of the dimeric complex [(η5‐C5Me5)RhCl2]2 with 3‐phenyl‐2H‐azirine (az) in CH2Cl2 at room temperature in a 1:2 molar ratio afforded the neutral mono‐azirine complex [(η5‐C5Me5)RhCl2(az)]. The subsequent reaction of [(η5‐C5Me5)RhCl2]2 with six equivalents of az and 4 equivalents of AgOTf yielded the cationic tris‐azirine complex [(η5‐C5Me5)Rh(az)3](OTf)2. After purification, all complexes have been fully characterized. The molecular structures of the novel rhodium(III ) complexes exhibit slightly distorted octahedral coordination geometries around the metal atoms.  相似文献   

16.
B. Khera  N.K. Kaushik 《Polyhedron》1984,3(5):611-613
A series of (C9H7)2Zr(OAr)Cl and (C9H7)2Zr(OAr)2 complexes, where Ar = C6H5, p-ClC6H4, α-C10H7, or β-C10H7, have been synthesised by the reaction of bis(indenyl)zirconium(IV)-dichloride with an appropriate phenol in a 1:1 and 1:2 molar ratio in refluxing benzene in the presence of triethylamine. These complexes have been characterised by elemental analyses, conductance measurements and spectral (IR, 1H NMR and electronic) studies.  相似文献   

17.
Vlad M. Iluc 《Tetrahedron》2006,62(32):7577-7582
Addition of hydridosilanes, Ar2SiHX, to the labile Ni(0) benzene complex [(dtbpe)Ni]2(C6H6) (1; dtbpe=1,2-bis(di-tert-butylphosphino)ethane) gives mononuclear Ni(II) hydride silyl complexes of the formulation (dtbpe)Ni(μ-H)SiAr2X (2, X=H, Ar=Mes; 3, X=H, Ar=Ph; 4, X=Me, Ar=Ph; 5, X=Cl, Ar=Ph). Although the crystal structures of two representatives of the series indicate square-planar coordination around nickel, in solution structures having apparent C2v symmetry are observed. We propose that this behavior is due to a fluxional process that involves η2-SiH intermediates. Other data are also consistent with the facile reductive elimination of the silane to regenerate nickel(0) products. Oxidation of 2 and 3 with triphenylcarbenium tetrakis(pentafluorophenyl)borate results in silane elimination and formation of [(dtbpe)Ni(η3-C6H5CPh2)+][B(C6F5)4] (6), the structure of which shows the CPh3 ligand bound to a Ni(II) center through a phenyl ring in an η3-allylic fashion.  相似文献   

18.
The hydroxo complex (Bu4N)2[Ni2(C6F5)4(μ-OH)2]reacts with 2,3,4,5,6-pentafluoro benzenamine (C6F5-NH2), 1,3-diaryltriaz-1-enes (ArNH? N=N? Ar, Ar = Ph, 4-MeC6H4, 4-MeOC6H4), 7-aza-1H-indole (= 1H-pyrrolo[2.3-b]pyridine; Hazind), N-phenylpyridin-2-amine(pyNHPh), and N-phenylpyridine-2-carboxamide (py-CONHPh) at room temperature in acetone to give the binuclear complexes (Bu4N)2[Ni2(C6F5)4(μ-C6F5NH)2] ( 1 ) and (Bu4N)2[{Ni(C6F5)2} 2(μ-OH)(μ-azind)] ( 2 ) and the mononuclear complexes Bu4N[Ni(C6F5)2(ArN3Ar)] ( 3 – 5 ), Bu4N[Ni(C6F5)2(pyNPh)] ( 6 ), and Bu4N[Ni(C6F5)2(pyCONPh)] ( 7 ). The hydroxo.complex (Bu4N)2[{Ni(C6F5)2-(μ-OH)}2] promotes the nucleophilic addition of water to pyridine-2-carbonitrile, 2-aminoacetonitrile, and 2-(dimethylamino)acetonitrile, and complexes 8 – 10 containing pyridine-2-carboxamidato, 2-aminoacetamidato and 2-(dimethylamino)acetamidato ligands are formed. Analytical (C, H, N) and spectroscopic (IR, 1H and 19F-NMR, and FAB-MS) data were used for structural assignments. A single-crystal X-ray diffraction study of (Bu4N)2[{Ni(C6F5)2}2(μ-OH)(μ-azind)] ( 2 ) established the binuclear nature of the anion; the two Ni-atoms are bridged by an OH group and a 7-aza-7H-indol-7-yl group, but the central Ni? O? Ni? N? C? N ring is not planar, the dihedral angle between the Ni? O? Ni and Ni? N? C? N? Ni planes being 84.4°.  相似文献   

19.
Metal‐π‐Arene‐Interactions in the Solid‐State Structures of Two Lewis Donor‐Free Arylbis(cyclopentadienyl)lanthanoids Ar*Yb(C5H4Me)2 ( 1 ) and Ar*SmCp2 ( 2 ) (Ar* = 2,6‐Mes2C6H3) have been obtained by the reaction of LiAr* with Yb(C5H4Me)3 or SmCp3 in toluene. Red crystals of 1 and orange crystals of 2 were characterized by X‐ray structure analysis. The lanthanoids are η5‐coordinated to the cyclopentadienyl ligands and η1‐coordinated to the ipso carbon atom of the aryl groups. Additional π‐arene contacts to one mesityl group give rise to a different pyramidalisation of the metal centers, which depends on the size of the central lanthanoid atom.  相似文献   

20.
Pincer‐type palladium complexes are among the most active Heck catalysts. Due to their exceptionally high thermal stability and the fact that they contain PdII centers, controversial PdII/PdIV cycles have been often proposed as potential catalytic mechanisms. However, pincer‐type PdIV intermediates have never been experimentally observed, and computational studies to support the proposed PdII/PdIV mechanisms with pincer‐type catalysts have never been carried out. In this computational study the feasibility of potential catalytic cycles involving PdIV intermediates was explored. Density functional calculations were performed on experimentally applied aminophosphine‐, phosphine‐, and phosphite‐based pincer‐type Heck catalysts with styrene and phenyl bromide as substrates and (E)‐stilbene as coupling product. The potential‐energy surfaces were calculated in dimethylformamide (DMF) as solvent and demonstrate that PdII/PdIV mechanisms are thermally accessible and thus a true alternative to formation of palladium nanoparticles. Initial reaction steps of the lowest energy path of the catalytic cycle of the Heck reaction include dissociation of the chloride ligands from the neutral pincer complexes [{2,6‐C6H3(XPR2)2}Pd(Cl)] [X=NH, R=piperidinyl ( 1 a ); X=O, R=piperidinyl ( 1 b ); X=O, R=iPr ( 1 c ); X=CH2, R=iPr ( 1 d )] to yield cationic, three‐coordinate, T‐shaped 14e? palladium intermediates of type [{2,6‐C6H3(XPR2)2}Pd]+ ( 2 ). An alternative reaction path to generate complexes of type 2 (relevant for electron‐poor pincer complexes) includes initial coordination of styrene to 1 to yield styrene adducts [{2,6‐C6H3(XPR2)2}Pd(Cl)(CH2?CHPh)] ( 4 ) and consecutive dissociation of the chloride ligand to yield cationic square‐planar styrene complexes [{2,6‐C6H3(XPR2)2}Pd(CH2?CHPh)]+ ( 6 ) and styrene. Cationic styrene adducts of type 6 were additionally found to be the resting states of the catalytic reaction. However, oxidative addition of phenyl bromide to 2 result in pentacoordinate PdIV complexes of type [{2,6‐C6H3(XPR2)2}Pd(Br)(C6H5)]+ ( 11 ), which subsequently coordinate styrene (in trans position relative to the phenyl unit of the pincer cores) to yield hexacoordinate phenyl styrene complexes [{2,6‐C6H3(XPR2)2}Pd(Br)(C6H5)(CH2?CHPh)]+ ( 12 ). Migration of the phenyl ligand to the olefinic bond gives cationic, pentacoordinate phenylethenyl complexes [{2,6‐C6H3(XPR2)2}Pd(Br)(CHPhCH2Ph)]+ ( 13 ). Subsequent β‐hydride elimination induces direct HBr liberation to yield cationic, square‐planar (E)‐stilbene complexes with general formula [{2,6‐C6H3(XPR2)2}Pd(CHPh?CHPh)]+ ( 14 ). Subsequent liberation of (E)‐stilbene closes the catalytic cycle.  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