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.
Electrophilic terminal phosphinidene complexes [Ar‐Ar‐P‐W(CO)5] (Ar‐Ar: biaryl or an analogue thereof) undergo a spontaneous insertion of the phosphorus atom into the vicinal C H bonds to give annelated phospholes. Twelve examples are described, including biphenyl, thienyl, pyrrolyl, and benzofuryl groups as biaryl moieties. The activation energy of the insertion reaction is quite low (about 2 kcal mol−1).  相似文献   

2.
In the present communication we describe two examples of a new kind of configurationally stable non‐biaryl atropisomers in which the Ar‐N bond is the chiral axis, namely 1‐(o‐nitrophenyl)‐2‐aryl‐3‐methyl‐1,4,5,6‐tetrahydropyrimidinium iodides 1. Stereochemical features of such compounds are analyzed on the basis of their 1H and 13C one‐ and two‐dimensional nmr spectra. A comparison is made with the corresponding amidines 2 .  相似文献   

3.
The first example of NO insertion into a Bi?C bond has been found in the direct reaction of NO with a Bi3+ complex of the unusual (C6H2tBu2‐3,5‐O‐4)2? oxyaryl dianionic ligand, namely, Ar′Bi(C6H2tBu2‐3,5‐O‐4) [Ar′=2,6‐(Me2NCH2)2C6H3] ( 1 ). The oximate complexes [Ar′Bi(ONC6H2‐3,5‐tBu2‐4‐O)]2(μ‐O) ( 3 ) and Ar′Bi(ONC6H2‐3,5‐tBu2‐4‐O)2 ( 4 ) were formed as a mixture, but can be isolated in pure form by reaction of NO with a Bi3+ complex of the [O2C(C6H2tBu2‐3‐5‐O‐4]2? oxyarylcarboxy dianion, namely, Ar′Bi[O2C(C6H2tBu2‐3‐5‐O‐4)‐κ2O,O’]. Reaction of 1 with Ph3CSNO gave an oximate product with (Ph3CS)1? as an ancillary ligand, (Ph3CS)(Ar′)Bi(ONC6H2‐3,5‐tBu2‐4‐O) ( 5 ).  相似文献   

4.
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   

5.
In Suzuki–Miyaura reactions, anionic bases F? and OH? (used as is or generated from CO32? in water) play multiple antagonistic roles. Two are positive: 1) formation of trans‐[Pd(Ar)F(L)2] or trans‐[Pd(Ar)‐ (L)2(OH)] (L=PPh3) that react with Ar′B(OH)2 in the rate‐determining step (rds) transmetallation and 2) catalysis of the reductive elimination from intermediate trans‐[Pd(Ar)(Ar′)(L)2]. Two roles are negative: 1) formation of unreactive arylborates (or fluoroborates) and 2) complexation of the OH group of [Pd(Ar)(L)2(OH)] by the countercation of the base (Na+, Cs+, K+).  相似文献   

6.
The molybdenum‐catalyzed asymmetric ring‐closing metathesis of the various Cs‐symmetric (π‐arene)chromium substrates provides the corresponding bridged planar‐chiral (π‐arene)chromium complexes in excellent yields with up to >99 % ee. With a bulky and unsymmetrical substituent, such as N‐indolyl or 1‐naphthyl, at the 2‐positions of the η6‐1,3‐diisopropenylbenzene ligands, both biaryl‐based axial chirality and π‐arene‐based planar chirality are simultaneously induced in the products. The axial chirality is retained even after the removal of the dicarbonylchromium fragment, and the chiral biaryl/heterobiaryl compounds are obtained with complete retention of the enantiopurity.  相似文献   

7.
Three new three‐dimensional geometry bulky α‐diimine ligands ( L ) containing electron‐donating and electron‐withdrawing groups, 9,10‐dihydro‐9,10‐ethanoanthracene‐11,12‐di(Ar)imine (Ar = p‐PhCH3, L1 ; Ar=p‐PhCl, L2 ; Ar=p‐PhCF3, L3 .), and their corresponding single Ni(II) catalysts, NiL2Br2 ( Ni(L1)2Br2 , Ni(L2)2Br2 , and Ni(L3)2Br2 , were synthesized and the molecular structure were determined by X‐ray crystallography. All NiL2Br2 catalysts were tested for norbornene polymerization and copolymerization of norbornene with 1‐alkene after activation with B(C6F5)3. The results that the polymerization catalytic activities for norbornene up to 105 gpolymer/molNi·h even at 140 °C, shown that NiL2Br2 catalysts have high thermal stability. Meanwhile, catalysts with electron‐withdrawing groups could achieve higher reactivity. The obtained poly(NB‐co‐1‐alkene)s were confirmed to be vinyl‐addition copolymers and noncrystalline. All copolymers exhibited high 1‐alkenes insertion ratio, good thermal stability (Td > 375 °C), high molecular weight (up to 105 g/mol), good solubility in common organic solvents and could be processed into films with good transparency in the visible region. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 3495–3505  相似文献   

8.
Reaction conditions for the three‐component synthesis of aryl 1,3‐diketones are reported applying the palladium‐catalyzed carbonylative α‐arylation of ketones with aryl bromides. The optimal conditions were found by using a catalytic system derived from [Pd(dba)2] (dba=dibenzylideneacetone) as the palladium source and 1,3‐bis(diphenylphosphino)propane (DPPP) as the bidentate ligand. These transformations were run in the two‐chamber reactor, COware, applying only 1.5 equivalents of carbon monoxide generated from the CO‐releasing compound, 9‐methylfluorene‐9‐carbonyl chloride (COgen). The methodology proved adaptable to a wide variety of aryl and heteroaryl bromides leading to a diverse range of aryl 1,3‐diketones. A mechanistic investigation of this transformation relying on 31P and 13C NMR spectroscopy was undertaken to determine the possible catalytic pathway. Our results revealed that the combination of [Pd(dba)2] and DPPP was only reactive towards 4‐bromoanisole in the presence of the sodium enolate of propiophenone suggesting that a [Pd(dppp)(enolate)] anion was initially generated before the oxidative‐addition step. Subsequent CO insertion into an [Pd(Ar)(dppp)(enolate)] species provided the 1,3‐diketone. These results indicate that a catalytic cycle, different from the classical carbonylation mechanism proposed by Heck, is operating. To investigate the effect of the dba ligand, the Pd0 precursor, [Pd(η3‐1‐PhC3H4)(η5‐C5H5)], was examined. In the presence of DPPP, and in contrast to [Pd(dba)2], its oxidative addition with 4‐bromoanisole occurred smoothly providing the [PdBr(Ar)(dppp)] complex. After treatment with CO, the acyl complex [Pd(CO)Br(Ar)(dppp)] was generated, however, its treatment with the sodium enolate led exclusively to the acylated enol in high yield. Nevertheless, the carbonylative α‐arylation of 4‐bromoanisole with either catalytic or stoichiometric [Pd(η3‐1‐PhC3H4)(η5‐C5H5)] over a short reaction time, led to the 1,3‐diketone product. Because none of the acylated enol was detected, this implied that a similar mechanistic pathway is operating as that observed for the same transformation with [Pd(dba)2] as the Pd source.  相似文献   

9.
A series of 2‐aryl‐ and 2‐heteroaryl‐substituted 3,5‐dimethoxy‐1,4‐benzoquinones (compounds 27 – 36 ) have been synthesized by cross‐coupling of (2,3,4,6‐tetramethoxyphenyl)boronic acid ( 2 ) with aromatic bromides or iodides in the presence of [Pd0(Ph3)4] and Na2CO3, followed by AgO‐promoted oxidation of the resulting biaryl compounds 17 – 26 .  相似文献   

10.
The Buchwald–Hartwig amination of nitroarenes was achieved for the first time by using palladium catalysts bearing dialkyl(biaryl)phosphine ligands. These cross‐coupling reactions of nitroarenes with diarylamines, arylamines, and alkylamines afforded the corresponding substituted arylamines. A catalytic cycle involving the oxidative addition of the Ar−NO2 bond to palladium(0) followed by nitrite/amine exchange is proposed based on a stoichiometric reaction.  相似文献   

11.
From the kinetic data on the transmetalation/reductive elimination in fluoride‐promoted Hiyama reactions, obtained using electrochemical techniques, it has been established that fluoride ions play three roles. F? reacts with trans‐[ArPdBrL2] (L=PPh3) to form trans‐[ArPdFL2], which reacts with Ar′Si(OMe)3 in the rate‐determining transmetalation, whereas trans‐[ArPdBrL2] does not react with Ar′Si(OMe)3. F? reacts with Ar′Si(OMe)3 to deliver the unreactive silicate Ar′SiF(OMe)3?, thus leading to two antagonistic kinetic effects. In addition, F? catalyzes the reductive elimination from intermediate trans‐[ArPdAr′L2].  相似文献   

12.
Hydration has a drastic impact on the structure and function of flexible biomolecules, such as aromatic ethylamino neurotransmitters. The structure of monohydrated protonated phenylethylamine (H+PEA?H2O) is investigated by infrared photodissociation (IRPD) spectroscopy of cold cluster ions by using rare‐gas (Rg=Ne and Ar) tagging and dispersion‐corrected density functional theory calculations at the B3LYP‐D3/aug‐cc‐pVTZ level. Monohydration of this prototypical neurotransmitter gives an insight into the first step of the formation of its solvation shell, especially regarding the competition between intra‐ and intermolecular interactions. The spectra of Rg‐tagged H+PEA?H2O reveal the presence of a stable insertion structure in which the water molecule is located between the positively charged ammonium group and the phenyl ring of H+PEA, acting both as a hydrogen bond acceptor (NH+???O) and donor (OH???π). Two other nearly equivalent isomers, in which water is externally H bonded to one of the free NH groups, are also identified. The balance between insertion and external hydration strongly depends on temperature.  相似文献   

13.
The mechanism of Stille reactions (cross‐coupling of ArX with Ar′SnnBu3) performed in the presence of fluoride ions is established. A triple role for fluoride ions is identified from kinetic data on the rate of the reactions of trans‐[ArPdBr(PPh3)2] (Ar=Ph, p‐(CN)C6H4) with Ar′SnBu3 (Ar′=2‐thiophenyl) in the presence of fluoride ions. Fluoride ions promote the rate‐determining transmetallation by formation of trans‐[ArPdF(PPh3)2], which reacts with Ar′SnBu3 (Ar′=Ph, 2‐thiophenyl) at room temperature, in contrast to trans‐[ArPdBr(PPh3)2], which is unreactive. However, the concentration ratio [F?]/[Ar′SnBu3] must not be too high, because of the formation of unreactive anionic stannate [Ar′Sn(F)Bu3]?. This rationalises the two kinetically antagonistic roles exerted by the fluoride ions that are observed experimentally, and is found to be in agreement with the kinetic law. In addition, fluoride ions promote reductive elimination from trans‐[ArPdAr′(PPh3)2] generated in the transmetallation step.  相似文献   

14.
A series of Al(III) and Sn(II) diiminophosphinate complexes have been synthesized. Reaction of Ph(ArCH2)P(?NBut)NHBut (Ar = Ph, 3 ; Ar = 8‐quinolyl, 4 ) with AlR3 (R = Me, Et) gave aluminum complexes [R2Al{(NBut)2P(Ph)(CH2Ar)}] (R = Me, Ar = Ph, 5 ; R = Me, Ar = 8‐quinolyl, 6 ; R = Et, Ar = Ph, 7 ; R = Et, Ar = quinolyl, 8 ). Lithiated 3 and 4 were treated with SnCl2 to afford tin(II) complexes [ClSn{(NBut)2P(Ph)(CH2Ar)}] (Ar = Ph, 9 ; Ar = 8‐quinolyl, 10 ). Complex 9 was converted to [(Me3Si)2NSn{(NBut)2P(Ph)(CH2Ph)}] ( 11 ) by treatment with LiN(SiMe3)2. Complex 11 was also obtained by reaction of 3 with [Sn{N(SiMe3)2}2]. Complex 9 reacted with [LiOC6H4But‐4] to yield [4‐ButC6H4OSn{(NBut)2P(Ph)(CH2Ph)}] ( 12 ). Compounds 3–12 were characterized by NMR spectroscopy and elemental analysis. The structures of complexes 6 , 10 , and 11 were further characterized by single crystal X‐ray diffraction techniques. The catalytic activity of complexes 5–8 , 11 , and 12 toward the ring‐opening polymerization of ε‐caprolactone (CL) was studied. In the presence of BzOH, the complexes catalyzed the ring‐opening polymerization of ε‐CL in the activity order of 5 > 7 ≈ 8 > 6 ? 11 > 12 , giving polymers with narrow molecular weight distributions. The kinetic studies showed a first‐order dependency on the monomer concentration in each case. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4621–4631, 2006  相似文献   

15.
Diaryl‐substituted triazenides Ar(Ar′)N3HgX [Ar/Ar′ = Dmp/Mph, X = Cl ( 2a ), Br ( 3a ), I ( 4a ); Ar/Ar′ = Dmp/Tph, X = Cl ( 2b ), I ( 4b ) with Mph = 2‐MesC6H4, Mes = 2,4,6‐Me3C6H2, Tph = 2′,4′,6′‐triisopropylbiphenyl‐2‐yl and Dmp = 2,6‐Mes2C6H3] were synthesized by salt‐metathesis reactions in ethyl ether from the readily available starting materials Ar(Ar′)N3Li and HgX2. These compounds may be used for redox‐transmetallation reactions with rare‐earth or alkaline earth metals. Thus, reaction of 4b or 2b with magnesium or ytterbium in tetrahydrofuran afforded the triazenides Dmp(Tph)N3MX(thf) ( 5b : M = Mg, X = I; 6b : M = Yb, X = Cl) in good yield. All new compounds were characterized by melting point, 1H and 13C NMR spectroscopy and for selected species by IR spectroscopy or mass spectrometry. In addition, the solid‐state structures of triazenides 2a , 2b , 3a , 4b , 5b and 6b were investigated by single‐crystal X‐ray diffraction.  相似文献   

16.
One‐electron reduction of C2‐arylated 1,3‐imidazoli(ni)um salts (IPrAr)Br (Ar=Ph, 3 a ; 4‐DMP, 3 b ; 4‐DMP=4‐Me2NC6H4) and (SIPrAr)I (Ar=Ph, 4 a ; 4‐Tol, 4 b ) derived from classical NHCs (IPr=:C{N(2,6‐iPr2C6H3)}2CHCH, 1 ; SIPr=:C{N(2,6‐iPr2C6H3)}2CH2CH2, 2 ) gave radicals [(IPrAr)]. (Ar=Ph, 5 a ; 4‐DMP, 5 b ) and [(SIPrAr)]. (Ar=Ph, 6 a ; 4‐Tol, 6 b ). Each of 5 a , b and 6 a , b exhibited a doublet EPR signal, a characteristic of monoradical species. The first solid‐state characterization of NHC‐derived carbon‐centered radicals 6 a , b by single‐crystal X‐ray diffraction is reported. DFT calculations indicate that the unpaired electron is mainly located at the original carbene carbon atom and stabilized by partial delocalization over the adjacent aryl group.  相似文献   

17.
An atropisomeric biaryl molecule with a given absolute configuration could present two opposite helical conformations through the rotation around C? C single bond. To the best of our knowledge, the biaryl system is the simplest helical inversion model apart from stereomutation between two enantiomers. Herein, we first report such true helical inversion phenomena of biaryl compounds. Two [MoVIO2(L)]‐type complexes, in which L is a tridentate dioxoanionic pyridine O,N,O‐ligand, are coalesced on the 2,2′,3,3′‐positions of an (R)‐1,1′‐binaphthyl unit and an intramolecular dioxo bridge is formed by two Mo?O???Mo interactions. Exterior strong donors can coordinate to molybdenum to interrupt this dioxo bridge and inversions from negative to positive chirality are explicitly observed by circular dichroism spectroscopy, consistent with single‐crystal X‐ray diffraction analyses.  相似文献   

18.
This work describes the synthesis and full characterization of elusive chloropnictenium ion salts of the type [RAr*N(SiMe)ECl][A] (RAr*=2,6‐(CHPh2)‐4‐R‐C6H2, R=Me, tBu; E=Sb, Bi; A?=GaCl4, Al(OCH(CF3)2)4). In these species the cation is significantly stabilized by weak arene interactions to flanking phenyl groups of the RAr* moiety. In this context the bonding situation has been studied by computational means and the reactivity towards the Lewis base 4‐dimethylaminopyridine (dmap) was investigated.  相似文献   

19.
A bulky substituted stannane Ar*SnH3 (Ar*=2,6‐(2′,4′,6′‐triisopropylphenyl)phenyl) was treated with the well‐known frustrated Lewis pair (FLP) Pt Bu3/B(C6F5)3 in varying stoichiometries. To some degree, hydride abstraction and adduct formation is observed, leading to [Ar*SnH2(Pt Bu3)]+ which is rather unreactive toward further dehydrogenation. In a competing process, the FLP proved to be capable of completely striping‐off hydrogen and hydrides to generate the first cationic phosphonio‐stannylene [Ar*Sn(Pt Bu3)]+. This behavior provides insight into the activation/abstraction mechanism processes involved in these Group 14 hydride derivatives.  相似文献   

20.
A series of Group 7 Fischer carbene complexes, [Cp(CO)2MnI=C(OEt)Ar] (Cp=cyclopentadienyl, Ar=Th=thienyl ( 1 a ), Ar=Fu=furyl ( 2 a ), Ar=Fc=ferrocenyl ( 3 a )) and biscarbene complexes, [Cp(CO)2Mn?C(OEt)?Ar′?(OEt)C?Mn(CO)2Cp] (Ar′=Th′=2,5‐thienylene ( 1 b ), Ar′=Fu′=2,5‐furylene ( 2 b ), Ar′=Fc′=1,1′‐ferrocendiyl ( 3 b )) was synthesized and characterized. Chemical oxidation of [Cp(CO)2Mn?C(OEt)Fc] ( 3 a ) and isolation of the oxidised species [3 a][PF6] possessing a MnII centre proved possible below ?30 °C in dichloromethane solution. The ESR spectrum of the transiently stable radical cation, [3 a][PF6] , confirmed the presence of a low‐spin MnII centre characterized by a rhombic g tensor (gx=1.975, gy=2.007 and gz=2.130) in frozen dichloromethane at 77 K with 55 Mn hyperfine coupling constants A1, A2 and A3 of 115, 33 and 43 G, respectively. Electrochemical studies demonstrated the influence of the Ar substituent on the oxidation potential. All complexes showed that the redox potentials of carbene double bond reduction and MnI oxidation were dependent on the type of Ar group, but only 3 b showed resolved oxidations for the two MnI centres. Surprisingly, MnI oxidation occurs at lower potentials than ferrocenyl oxidation. Density functional theory (DFT) calculations were carried out to delineate the nature of the species involved in the oxidation and reduction processes and clearly confirm that oxidation of MnI is favoured over that of ferrocene.  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