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 46 毫秒
1.
The tin(II) complexes {LOx}Sn(X) ({LOx}?=aminophenolate ancillary) containing amido ( 1 – 4 ), chloro ( 5 ), or lactyl ( 6 ) coligands (X) promote the ring‐opening polymerization (ROP) of cyclic esters. Complex 6 , which models the first insertion of L ‐lactide, initiates the living ROP of L ‐LA on its own, but the amido derivatives 1 – 4 require the addition of alcohol to do so. Upon addition of one to ten equivalents of iPrOH, precatalysts 1 – 4 promote the ROP of trimethylene carbonate (TMC); yet, hardly any activity is observed if tert‐butyl (R)‐lactate is used instead of iPrOH. Strong inhibition of the reactivity of TMC is also detected for the simultaneous copolymerization of L ‐LA and TMC, or for the block copolymerization of TMC after that of L ‐LA. Experimental and computational data for the {LOx}Sn(OR) complexes (OR=lactyl or lactidyl) replicating the active species during the tin(II)‐mediated ROP of L ‐LA demonstrate that the formation of a five‐membered chelate is largely favored over that of an eight‐membered one, and that it constitutes the resting state of the catalyst during this (co)polymerization. Comprehensive DFT calculations show that, out of the four possible monomer insertion sequences during simultaneous copolymerization of L ‐LA and TMC: 1) TMC then TMC, 2) TMC then L ‐LA, 3) L ‐LA then L ‐LA, and 4) L ‐LA then TMC, the first three are possible. By contrast, insertion of L ‐LA followed by that of TMC (i.e., insertion sequence 4) is endothermic by +1.1 kcal mol?1, which compares unfavorably with consecutive insertions of two L ‐LA units (i.e., insertion sequence 3) (?10.2 kcal mol?1). The copolymerization of L ‐LA and TMC thus proceeds under thermodynamic control.  相似文献   

2.
The factors governing the stability and the reactivity towards cyclic esters of heteroleptic complexes of the large alkaline earth metals (Ae) have been probed. The synthesis and stability of a family of heteroleptic silylamido and alkoxide complexes of calcium [{LOi}Ca? Nu(thf)n] supported by mono‐anionic amino ether phenolate ligands (i=1, {LO1}?=4‐(tert‐butyl)‐2,6‐bis(morpholinomethyl)phenolate, Nu?=N(SiMe2H)2?, n=0, 4 ; i=2, {LO2}?=2,4‐di‐tert‐butyl‐6‐{[2‐(methoxymethyl)pyrrolidin‐1‐yl]methyl}phenolate, Nu?=N(SiMe2H)2?, n=0, 5 ; i=4, {LO4}?=2‐{[bis(2‐methoxyethyl)amino]methyl}‐4,6‐di‐tert‐butylphenolate, Nu?=N(SiMe2H)2?, n=1, 6 ; Nu?=HC?CCH2O?, n=0, 7 ) and those of the related [{LO3}Ae? N(SiMe2H)2] ({LO3}?=2‐[(1,4,7,10‐tetraoxa‐13‐azacyclopentadecan‐13‐yl)methyl]‐4,6‐di‐tert‐butylphenolate Ae=Ca, 1 ; Sr, 2 ; Ba, 3 ) have been investigated. The molecular structures of 1 , 2 , [( 4 )2], 6 , and [( 7 )2] have been determined by X‐ray diffraction. These highlight Ae???H? Si internal β‐agostic interactions, which play a key role in the stabilization of [{LOi}Ae? N(SiMe2H)2] complexes against ligand redistribution reactions, in contrast to regular [{LOi}Ae? N(SiMe3)2]. Pulse‐gradient spin‐echo (PGSE) NMR measurements showed that 1 , 4 , 6 , and 7 are monomeric in solution. Complexes 1 – 7 mediate the ring‐opening polymerization (ROP) of L ‐lactide highly efficiently, converting up to 5000 equivalents of monomer at 25 °C in a controlled fashion. In the immortal ROP performed with up to 100 equivalents of exogenous 9‐anthracenylmethanol or benzyl or propargyl alcohols as a transfer agent, the activity of the catalyst increased with the size of the metal ( 1 < 2 < 3 ). For Ca‐based complexes, the enhanced electron‐donating ability of the ancillary ligand favored catalyst activity ( 1 > 6 > 4 ≈ 5 ). The nature of the alcohol had little effect over the activity of the binary catalyst system 1 /ROH; in all cases, both the control and end‐group fidelity were excellent. In the living ROP of L ‐LA, the HC?CCH2O? initiating group (as in 7 ) proved superior to N(SiMe2H)2? or N(SiMe3)2? (as in 6 or [{LO4}Ca? N(SiMe3)2] ( B ), respectively).  相似文献   

3.
The tetravalent germanium and tin compounds of the general formulae Ph*EX3 (Ph* = C6H3Trip‐2,6, Trip = C6H2iPr3‐2,4,6; E = Sn, X = Cl ( 1a ), Br ( 1b ); E = Ge, X = Cl ( 2 )) are synthesized by reaction of Ph*Li·OEt2 with EX4. The subsequent reaction of 1a , b with LiP(SiMe3)2 leads to Ph*EP(SiMe3)2 (E = Sn ( 3 ), Ge ( 4 )) and the diphosphane (Me3Si)2PP(SiMe3)2 by a redox reaction. In an alternative approach 3 and 4 are synthesized by using the corresponding divalent compounds Ph*ECl (E = Ge, Sn) in the reaction with LiP(SiMe3)2. The reactivity of Ph*SnCl is extensively investigated to give with LiP(H)Trip a tin(II)‐phosphane derivative Ph*SnP(H)Trip ( 6 ) and with Li2PTrip a proposed product [Ph*SnPTrip] ( 7 ) with multiple bonding between tin and phosphorus. The latter feature is confirmed by DFT calculations on a model compound [PhSnPPh]. The reaction with Li[H2PW(CO)5] gives the oxo‐bridged tin compound [Ph*Sn{W(CO)5}(μ‐O)2SnPh*] ( 8 ) as the only isolable product. However, the existence of 8 as the bis‐hydroxo derivative [Ph*Sn{W(CO)5}(μ‐OH)2SnPh*] ( 8a ) is also possible. The SnIV derivatives Ph*Sn(OSiMe3)2Cl ( 9 ) and [Ph*Sn(μ‐O)Cl]2 ( 10 ) are obtained by the oxidation of Ph*SnCl with bis(trimethylsilyl)peroxide and with Me3NO, respectively. Besides the spectroscopic characterization of the isolated products compounds 1a , 2 , 3 , 4 , 8 , and 10 are additionally characterized by X‐ray diffraction analysis.  相似文献   

4.
Polyol Metal Complexes.471) Crystalline D ‐Mannose‐Copper Complexes from Fehling Solutions Blue, unstable crystals of K3[Cu5(β‐D ‐Manp)4H—13] · α‐D ‐Manp · 16.5 H2O ( 1 ), which contain a pentanuclear cupric complex of the reducing sugar D ‐mannose in its β‐pyranose form (β‐D ‐Manp), have been obtained from ice‐cold aqueous alkaline solutions. The homoleptic pentacuprate contains bridging mannopyranose ligands, which are charged 4— and 2.5—. Addition of ethylenediamine (en) to such Fehling solutions yields N, N′‐Bis(β‐D ‐mannopyranosyl)‐ethylenediamine (L) as a condensation product of the diamine and mannopyranose. Crystals of [(en)2Cu7(β‐D ‐Manp1, 2, 3, 4H—4)2(L2, 3, 4H—3)2] · 26.6 H2O ( 2 ) could be isolated. The heptanuclear cupric complex is a structural derivative of the homoleptic mannose complex.  相似文献   

5.
Synthesis and Characterization of InIII–SnII‐Halogenido‐Alkoxides and of Indiumtri‐ tert ‐butoxide Through sodium halide elimination between Indium(III) halides and sodium‐tri‐tert‐butoxistannate(II) or sodium‐tri‐tert‐butoxigermanate(II) the three new heterometallic and heteroleptic alkoxo compounds THF · Cl2In(OtBu)3Sn ( 1 ), THF · Br2In(OtBu)3Sn ( 2 ), and THF · Cl2In‐ (OtBu)3Ge ( 3 ), have been synthesized. The molecular structures of 1 and 2 in the solid state follow from single crystal X‐ray structure determinations while structural changes in solution may be derived from temperature dependant NMR spectroscopy. The crystal structures of compounds 1 and 2 are despite different halide atoms isostructural. Both crystallize in the ortho‐rhombic crystal system in space group Pbca with eight molecules per unit cell. The heavy atoms occupy the apical positions of empty trigonal bipyramids of almost point symmetry Cs(m) and are connected through oxygen atoms occupying the equatorial positions. The indium atoms in both compounds are in the centers of distorted octahedra from 4 oxygen and 2 halogen atoms whereas the tin atoms are coordinated by three oxygen atoms in a trigonal pyramidal fashion. Although the coordinative bonding of THF to indium leads to an asymmetry of the molecule the NMR spectra in solution are simple showing a more complex pattern at lower temperatures. Tri(tert‐butoxi)indium [In(OtBu)3]2 ( 4 ), is obtained through alcoholysis of In(N(Si(CH3)3)2)3 using tert‐butanol in toluene and is crystallized from hexane. The X‐ray structure determination of 4 seems to be the first one of a homoleptic and homometallic indiumalkoxide. Compound 4 crystallizes in the monoclinic crystal system in a dimeric form with eight molecules in the unit cell of space group C2/c. The dimeric units have C2 symmetry and an almost planar In2O2 ring which originates from oxygen bridging of the monomers. Through this mutual Lewis acid base interaction the indium atoms get four oxygen ligands in a distorted tetrahedral environment.  相似文献   

6.
The heteroleptic neutral tri‐tert‐butoxysilanethiolate of cobalt(II) incorporating ammonia as additional ligand ( 1 ) has been prepared by the reaction of a cobalt(II) ammine complex with tri‐tert‐butoxysilanethiol in water. Complex 1 , dissolved in hexane, undergoes oxidation in an ammonia saturated atmosphere to the ionic cobalt(III) compound 2 . Molecular and crystal structures of 1 and 2 have been determined by single crystal X‐ray structural analysis. 1 forms a dimeric molecule [Co{μ‐SSi(OBut)3}{SSi(OBut)3}(NH3)]2 with a folded central Co2S2 ring and distorted tetrahedral ligand arrangement at both CoII atoms (CoNS3 core). The product 2 is composed of the octahedral CoIII complex cation [Co{SSi(OBut)3}2(NH3)4]+ and the tri‐tert‐butoxysilanethiolate anion. Within the crystal two pairs of ions interact by hydrogen bonds forming well separated entities. 1 and 2 are the first structurally characterized cobalt thiolates where metal is also bonded to ammonia and 2 is the first cobalt(III) silanethiolate.  相似文献   

7.
Structure and magnetic properties of N‐diisopropoxyphosphorylthiobenzamide PhC(S)‐N(H)‐P(O)(OiPr)2 ( HLI ) and N‐diisopropoxyphosphoryl‐N′‐phenylthiocarbamide PhN(H)‐C(S)‐N(H)‐P(O)(OiPr)2 ( HLII ) complexes with the CoII cation of formulas [Co{PhC(S)‐N‐P(O)(OiPr)2}2] ( 1 ), [Co{PhN(H)‐C(S)‐N‐P(O)(OiPr)2}2] ( 2 ), [Co{PhC(S)‐N(H)‐P(O)(OiPr)2}2{PhC(S)‐N‐P(O)(OiPr)2}2] ( 1a ) and [Co{PhC(S)‐N‐P(O)(OiPr)2}2}(2,2′‐bipy)] ( 3 ), [Co{PhC(S)‐N‐P(O)(OiPr)2}2(1,10‐phen)] ( 4 ), [Co{PhN(H)‐C(S)‐N‐P(O)(OiPr)2}2(2,2′‐bipy)] ( 5 ), [Co{PhN(H)‐C(S)‐N‐P(O)(OiPr)2}2(1,10‐phen)] ( 6 ) were investigated. Paramagnetic shifts in the 1H NMR spectrum were observed for high‐spin CoII complexes with HLI,II , incorporating the S‐C‐N‐P‐O chelate moiety and two aromatic chelate ligands. Investigation of the thermal dependence of the magnetic susceptibility has shown that the extended materials 1‐2 and 6 show ferromagnetic exchange between distorted tetrahedral ( 1 , 2 ) or octahedral ( 1a , 6 ) metal atoms whereas 3 and 5 show antiferromagnetic properties. Compound 4 behaves as a spin‐canted ferromagnet, an antiferromagnetic ordering taking place below a critical temperature, Tc = 115 K. Complexes 1 and 1a were investigated by single crystal X‐ray diffraction. The cobalt(II) atom in complex 1 resides a distorted tetrahedral O2S2 environment formed by the C=S sulfur atoms and the P=O oxygen atoms of two deprotonated ligands. Complex 1a has a tetragonal‐bipyramidal structure, Co(Oax)2(Oeq)2(Seq)2, and two neutral ligand molecules are coordinated in the axial positions through the oxygen atoms of the P=O groups. The base of the bipyramid is formed by two anionic ligands in the typical 1,5‐O,S coordination mode. The ligands are in a trans configuration.  相似文献   

8.
An improved synthetic route to homoleptic complex [Pt(CAACMe)2] (CAAC=cyclic (alkyl)(amino)carbenes) and convenient routes to new heteroleptic complexes of the form [Pt(CAACMe)(PR3)] are presented. Although the homoleptic complex was found to be inert to many reagents, oxidative addition and metal‐only Lewis pair (MOLP) formation was observed from one of the heteroleptic complexes. The spectroscopic, structural, and electrochemical properties of the zero‐valent complexes were explored in concert with density functional theory (DFT) and time‐dependent density functional theory (TD‐DFT) calculations. The homoleptic [Pt(CAAC)2] and heteroleptic [Pt(CAAC)(PR3)] complexes were found to be similar in their spectroscopic and structural properties, but their electrochemical behavior and reactivity differ greatly. The unusually strong color of the CAAC‐containing Pt0 complexes was investigated by TD‐DFT calculations and attributed to excitations into the LUMOs of the complexes, which are predominantly composed of bonding π interactions between Pt and the CAAC carbon atoms.  相似文献   

9.
Copper(I) halides with triphenyl phosphine and imidaozlidine‐2‐thiones (L ‐NMe, L ‐NEt, and L ‐NPh) in acetonitrile/methanol (or dichloromethane) yielded copper(I) mixed‐ligand complexes: mononuclear, namely, [CuCl(κ1‐S‐L ‐NMe)(PPh3)2] ( 1 ), [CuBr(κ1‐S‐L ‐NMe)(PPh3)2] ( 2 ), [CuBr(κ1‐S‐L ‐NEt)(PPh3)2] ( 5 ), [CuI(κ1‐S‐L ‐NEt)(PPh3)2] ( 6 ), [CuCl(κ1‐S‐L ‐NPh)(PPh3)2] ( 7 ), and [CuBr(κ1‐S‐L ‐NPh)(PPh3)2] ( 8 ), and dinuclear, [Cu21‐I)2(μ‐S‐L ‐NMe)2(PPh3)2] ( 3 ) and [Cu2(μ‐Cl)21‐S‐L ‐NEt)2(PPh3)2] ( 4 ). All complexes were characterized with analytical data, IR and NMR spectroscopy, and X‐ray crystallography. Complexes 2 – 4 , 7 , and 8 each formed crystals in the triclinic system with P$\bar{1}$ space group, whereas complexes 1 , 5 , and 6 crystallized in the monoclinic crystal system with space groups P21/c, C2/c, and P21/n, respectively. Complex 2 has shown two independent molecules, [(CuBr(κ1‐S‐L ‐NMe)(PPh3)2] and [CuBr(PPh3)2] in the unit cell. For X = Cl, the thio‐ligand bonded to metal as terminal in complex 4 , whereas for X = I it is sulfur‐bridged in complex 3 .  相似文献   

10.
Intramolecularly‐stabilized germanium, tin, and lead alkoxides of the type M(OCH2CH2NR2)2 [R = Et, M = Ge ( 1 ); R = Me, M = Sn ( 2 ); R = Me, M = Pb ( 3 )] are suitable precursors for the synthesis of group 14 chalcogenides ME (M = Ge, Sn, Pb; E = S, Se, Te) in scrambling reactions with (Me3Si)2S and (Et3Si)2E (E = Se, Te) at moderate temperatures via hot injection method. The reactions proceed with elimination of the corresponding silylether as was proven by in situ 1H NMR spectroscopy. The solid‐state structures of the homoleptic complex 1 and the heteroleptic complex ClGe(OC2H4NEt2) ( 4 ) were determined by single‐crystal X‐ray diffraction, whereas the group 14 chalcogenides were characterized by XRD, SEM, and EDX.  相似文献   

11.
A new CeIV complex [Ce{NH(CH2CH2N=CHC6H2‐3,5‐(tBu)2‐2‐O)2}(NO3)2] ( 1 ), bearing a dianionic pentadentate ligand with an N3O2 donor set, has been prepared by treating (NH4)2Ce(NO3)6 with the sodium salt of ligand L1 . Complex 1 in the presence of TEMPO and 4 Å molecular sieves (MS4 A) has been found to serve as a catalyst for the oxidation of arylmethanols using dioxygen as an oxidant. We propose an oxidation mechanism based on the isolation and reactivity study of a trivalent cerium complex [Ce{NH(CH2CH2N=CHC6H2‐3,5‐(tBu)2‐2‐O)2}(NO3)(THF)] ( 2 ), its side‐on μ‐O2 adduct [Ce{NH(CH2CH2N=CHC6H2‐3,5‐(tBu)2‐2‐O)2}(NO3)]2(μ‐η22‐O2) ( 3 ), and the hydroxo‐bridged CeIV complex [Ce{NH(CH2CH2N=CHC6H2‐3,5‐(tBu)2‐2‐O)2}(NO3)]2(μ‐OH)2 ( 4 ) as key intermediates during the catalytic cycle. Complex 2 was synthesized by reduction of 1 with 2,5‐dimethyl‐1,4‐bis(trimethylsilyl)‐1,4‐diazacyclohexadiene. Bubbling O2 into a solution of 2 resulted in formation of the peroxo complex 3 . This provides the first direct evidence for cerium‐catalyzed oxidation of alcohols under an O2 atmosphere.  相似文献   

12.
Two types of three‐arm and four‐arm, star‐shaped poly(D,L ‐lactic acid‐alt‐glycolic acid)‐b‐poly(L ‐lactic acid) (D,L ‐PLGA50‐b‐PLLA) were successfully synthesized via the sequential ring‐opening polymerization of D,L ‐3‐methylglycolide (MG) and L ‐lactide (L ‐LA) with a multifunctional initiator, such as trimethylolpropane and pentaerythritol, and stannous octoate (SnOct2) as a catalyst. Star‐shaped, hydroxy‐terminated poly(D,L ‐lactic acid‐alt‐glycolic acid) (D,L ‐PLGA50) obtained from the polymerization of MG was used as a macroinitiator to initiate the block polymerization of L ‐LA with the SnOct2 catalyst in bulk at 130 °C. For the polymerization of L ‐LA with the three‐arm, star‐shaped D,L ‐PLGA50 macroinitiator (number‐average molecular weight = 6800) and the SnOct2 catalyst, the molecular weight of the resulting D,L ‐PLGA50‐b‐PLLA polymer linearly increased from 12,600 to 27,400 with the increasing molar ratio (1:1 to 3:1) of L ‐LA to MG, and the molecular weight distribution was rather narrow (weight‐average molecular weight/number‐average molecular weight = 1.09–1.15). The 1H NMR spectrum of the D,L ‐PLGA50‐b‐PLLA block copolymer showed that the molecular weight and unit composition of the block copolymer were controlled by the molar ratio of L ‐LA to the macroinitiator. The 13C NMR spectrum of the block copolymer clearly showed its diblock structures, that is, D,L ‐PLGA50 as the first block and poly(L ‐lactic acid) as the second block. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 409–415, 2002  相似文献   

13.
Syntheses, Structure and Reactivity of η3‐1,2‐Diphosphaallyl Complexes and [{(η5‐C5H5)(CO)2W–Co(CO)3}{μ‐AsCH(SiMe3)2}(μ‐CO)] Reaction of ClP=C(SiMe2iPr)2 ( 3 ) with Na[Mo(CO)35‐C5H5)] afforded the phosphavinylidene complex [(η5‐C5H5)(CO)2Mo=P=C(SiMe2iPr)2] ( 4 ) which in situ was converted into the η1‐1,2‐diphosphaallyl complex [η5‐(C5H5)(CO)2Mo{η3tBuPPC(SiMe2iPr)2] ( 6 ) by treatment with the phosphaalkene tBuP=C(NMe2)2. The chloroarsanyl complexes [(η5‐C5H5)(CO)3M–As(Cl)CH(SiMe3)2] [where M = Mo ( 9 ); M = W ( 10 )] resulted from the reaction of Na[M(CO)35‐C5H5)] (M = Mo, W) with Cl2AsCH(SiMe3)2. The tungsten derivative 10 and Na[Co(CO)4] underwent reaction to give the dinuclear μ‐arsinidene complex [(η5‐C5H5)(CO)2W–Co(CO)3{μ‐AsCH(SiMe3)2}(μ‐CO)] ( 11 ). Treatment of [(η5‐C5H5)(CO)2Mo{η3tBuPPC(SiMe3)2}] ( 1 ) with an equimolar amount of ethereal HBF4 gave rise to a 85/15 mixture of the saline complexes [(η5‐C5H5)(CO)2Mo{η2tBu(H)P–P(F)CH(SiMe3)2}]BF4 ( 18 ) and [Cp(CO)2Mo{F2PCH(SiMe3)2}(tBuPH2)]BF4 ( 19 ) by HF‐addition to the PC bond of the η3‐diphosphaallyl ligand and subsequent protonation ( 18 ) and/or scission of the PP bond by the acid ( 19 ). Consistently 19 was the sole product when 1 was allowed to react with an excess of ethereal HBF4. The products 6 , 9 , 10 , 11 , 18 and 19 were characterized by means of spectroscopy (IR, 1H‐, 13C{1H}‐, 31P{1H}‐NMR, MS). Moreover, the molecular structures of 6 , 11 and 18 were determined by X‐ray diffraction analysis.  相似文献   

14.
A novel tridentate anilido‐aldimine ligand, [o‐C6H4(NHAr)? HC?NCH2CH2NMe2] (Ar = 2,6‐iPr2C6H3, L ‐H, 1 ), has been prepared by the condensation of N, N‐dimethylethylenediamine with one molar equivalent of 2‐fluoro‐benzaldehyde in hexane, followed by the addition of the lithium salt of diisopropylaniline in THF. Magnesium (Mg) and zinc (Zn) complexes supported by the tridentate anilido‐aldimine ligand have been synthesized and structurally characterized. Reaction of L ‐H ( 1 ) with an equivalent amount of MgnBu2 or ZnEt2 produces the monomeric complex [ L MgnBu] ( 2 ) or [ L ZnEt] ( 3 ), respectively. Experimental results show that complexes 2 and 3 are efficient catalysts for ring‐opening polymerization of ε‐caprolactone (CL) and L ‐lactide (LA) in the presence of benzyl alcohol and catalyze the polymerization of ε‐CL and L ‐LA in a controlled fashion yielding polymers with a narrow polydispersity index. In both polymerizations, the activity of Mg complex 2 is higher than that of Zn complex 3 , which is probably due to the higher Lewis acidity and better oxophilic nature of Mg2+ metal. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 4927–4936, 2009  相似文献   

15.
The scaffold geometries, stability and magnetic features of the (pyridine‐2‐yl)methanolate (L) supported wheel‐shaped transition‐metal complexes with compositions [M6L12] ( 1 ), [Na?(ML2)6]+ ( 2 ), and [M′?(ML2)6]2+ ( 3 ), in which M=CoII, NiII, CuII, and ZnII were investigated with density functional theory (DFT). The goals of this study are manifold: 1) To advance understanding of the magnetism in the synthesized compounds [Na?(ML2)6]+ and [M′?(ML2)6]2+ that were described in Angew. Chem. Int. Ed.­ 2010 , 49, 4443 ( I ‐{Na?Ni6}, I ‐{Ni′?Ni6}) and Dalton Trans.­ 2011 , 40, 10526 ( II ‐{Na?Co6}, II ‐{Co′?Co6}); 2) To disclose how the structural, electronic, and magnetic characteristics of 1 , 2 , and 3 change upon varying MII from d7 (Co2+) to d10 (Zn2+); 3) To estimate the influence of the Na+ and M′2+ ions (XQ+) occupying the central voids of 2 and 3 on the external and internal magnetic coupling interactions in these spin structures; 4) To assess the relative structural and electrochemical stabilities of 1 , 2 , and 3 . In particular, we focus here on the net spin polarization, the determination of the strength and the sign of the exchange coupling energies, the rationalization of the nature of the magnetic coupling, and the ground‐state structures of 1 , 2 , and 3 . Our study combines the broken symmetry DFT approach and the model Hamiltonian methodology implemented in the computational framework CONDON 2.0 for the modeling of molecular spin structures, to interpret magnetic susceptibility measurements of I ‐{Na?Ni6} and I ‐{Ni′?Ni6}. We illustrate that whereas the structures, stability and magnetism of 1 , 2 , and 3 are indeed influenced by the nature of 3d transition‐metals in the {M6} rims, the XQ+ ions in the inner cavities of 2 and 3 impact these properties to an even larger degree. As exemplified by I ‐{Ni′?Ni6}, such heptanuclear complexes exhibit ground‐state multiplets that cannot be described by simplistic model of spin‐up and spin‐down metal centers. Furthermore, we assess how future low‐temperature susceptibility measurements at high magnetic fields can augment the investigation of compound 3 with M=Co, Ni.  相似文献   

16.
The first solvent‐free cationic complexes of the divalent rare‐earth metals, [{RO}REII]+[A]? (REII=YbII, 1 ; EuII, 2 ) and [{LO}REII]+[A]? ([A]?=[H2N{B(C6F5)3}2]?; REII=YbII, 3 ; EuII, 4 ), have been prepared by using highly chelating monoanionic aminoether‐fluoroalkoxide ({RO}?) and aminoether‐phenolate ({LO}?) ligands. Complexes 1 and 2 are structurally related to their alkaline‐earth analogues [{RO}AE]+[A]? (AE=Ca, 5 ; Sr, 6 ). Yet, the two families behave very differently during catalysis of the ring‐opening polymerization (ROP) of L ‐lactide (L ‐LA) and trimethylene carbonate (TMC) performed under immortal conditions with excess BnOH as an exogenous chain‐transfer agent. The ligand was found to strongly influence the behavior of the REII complexes during ROP catalysis. The fluoroalkoxide REII catalysts 1 and 2 are not oxidized under ROP conditions, and compare very favorably with their Ca and Sr congeners 5 and 6 in terms of activity (turnover frequency (TOF) in the range 200–400 molL‐LA (molEu h?1)) and control over the parameters during the immortal ROP of L ‐LA (Mn,theorMn,SEC, Mw/Mn<1.05). The EuII‐phenolate 4 provided one of the most effective ROP cationic systems known to date for L ‐LA polymerization, exhibiting high activity (TOF up to 1 880 molL‐LA?(molEu h)?1) and good control (Mw/Mn=1.05). By contrast, upon addition of L ‐LA the YbII‐phenolate 3 immediately oxidizes to inactive REIII species. Yet, the cyclic carbonate TMC was rapidly polymerized by combinations of 3 (or even 1 ) and BnOH, revealing excellent activities (TOF=5000–7000 molTMC?(molEu h)?1) and unusually high control (Mn,theorMn,SEC, Mw/Mn<1.09); under identical conditions, the calcium derivative 5 was entirely inert toward TMC. Based on experimental and kinetic data, a new ligand‐assisted activated monomer ROP mechanism is suggested, in which the so‐called ancillary ligand plays a crucial role in the catalytic cycle. A coherent reaction pathway computed by DFT, compatible with the experimental data, supports the proposed scenario.  相似文献   

17.
Seven magnesium complexes ( 1–7 ) were synthesized by reaction of new ( L 3 ‐H – L 5 ‐H ) and previously reported ketoimine pro‐ligands with dibutyl magnesium and were isolated in 59–70% yields. Complexes 1–7 were characterized fully and consisted of bis‐ligated homoleptic ketoiminates coordinated in distorted octahedral geometry around the magnesium centers. The complexes were investigated for their ability to initiate the ring opening polymerization (ROP) of l ‐lactide (L‐LA) to poly‐lactic acid (PLA) and ?‐caprolactone (?CL) to poly‐caprolactone in the presence of 4‐fluorophenol co‐catalyst. For L‐LA polymerization, complexes containing ligand electron‐donating groups ( 1–5 ) achieved >90% conversion in 2 h at 100 °C, while the presence of CF3 groups in 6 and 7 slowed or resulted in no PLA detected. With ?CL, ROP initiated with 1–7 resulted in lower percentage conversion with similar electronic effects. Moderate molecular weight PLA polymeric material (14.3–21.3 kDa) with low polydispersity index values (1.23–1.56) was obtained, and ROP appeared to be living in nature. Copolymerization of L‐LA and ?CL yielded block copolymers only from the sequential polymerization of ?CL followed by L‐LA and not the reverse sequence of monomers or the simultaneous presence of both monomers. Polymers and copolymers were characterized with NMR, gel permeation chromatography, and differential scanning calorimetry. © 2018 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2019 , 57, 48–59  相似文献   

18.
Achiral P‐donor pincer‐aryl ruthenium complexes ([RuCl(PCP)(PPh3)]) 4c , d were synthesized via transcyclometalation reactions by mixing equivalent amounts of [1,3‐phenylenebis(methylene)]bis[diisopropylphosphine] ( 2c ) or [1,3‐phenylenebis(methylene)]bis[diphenylphosphine] ( 2d ) and the N‐donor pincer‐aryl complex [RuCl{2,6‐(Me2NCH2)2C6H3}(PPh3)], ( 3 ; Scheme 2). The same synthetic procedure was successfully applied for the preparation of novel chiral P‐donor pincer‐aryl ruthenium complexes [RuCl(P*CP*)(PPh3)] 4a , b by reacting P‐stereogenic pincer‐arenes (S,S)‐[1,3‐phenylenebis(methylene)]bis[(alkyl)(phenyl)phosphines] 2a , b (alkyl=iPr or tBu, P*CHP*) and the complex [RuCl{2,6‐(Me2NCH2)2C6H3}(PPh3)], ( 3 ; Scheme 3). The crystal structures of achiral [RuCl(equation/tex2gif-sup-3.gifPCP)(PPh3)] 4c and of chiral (S,S)‐[RuCl(equation/tex2gif-sup-6.gifPCP)(PPh3)] 4a were determined by X‐ray diffraction (Fig. 3). Achiral [RuCl(PCP)(PPh3)] complexes and chiral [RuCl(P*CP*)(PPh3)] complexes were tested as catalyst in the H‐transfer reduction of acetophenone with propan‐2‐ol. With the chiral complexes, a modest enantioselectivity was obtained.  相似文献   

19.
Methoxy‐modified β‐diimines HL 1 and HL 2 reacted with Y(CH2SiMe3)3(THF)2 to afford the corresponding bis(alkyl)s [L1Y(CH2SiMe3)2] ( 1 ) and [L2Y(CH2SiMe3)2] ( 2 ), respectively. Amination of 1 with 2,6‐diisopropyl aniline gave the bis(amido) counterpart [L1Y{N(H)(2,6‐iPr2? C6H3)}2] ( 3 ), selectively. Treatment of Y(CH2SiMe3)3(THF)2 with methoxy‐modified anilido imine HL 3 yielded bis(alkyl) complex [L3Y(CH2SiMe3)2(THF)] ( 4 ) that sequentially reacted with 2,6‐diisopropyl aniline to give the bis(amido) analogue [L3Y{N(H)(2,6‐iPr2? C6H3)}2] ( 5 ). Complex 2 was “base‐free” monomer, in which the tetradentate β‐diiminato ligand was meridional with the two alkyl species locating above and below it, generating tetragonal bipyramidal core about the metal center. Complex 3 was asymmetric monomer containing trigonal bipyramidal core with trans‐arrangement of the amido ligands. In contrast, the two cis‐located alkyl species in complex 4 were endo and exo towards the O,N,N tridentate anilido‐imido moiety. The bis(amido) complex 5 was confirmed to be structural analogue to 4 albeit without THF coordination. All these yttrium complexes are highly active initiators for the ring‐opening polymerization of L ‐LA at room temperature. The catalytic activity of the complexes and their “single‐site” or “double‐site” behavior depend on the ligand framework and the geometry of the alkyl (amido) species in the corresponding complexes. © 2007 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 5662–5672, 2007  相似文献   

20.
Insertion and Substitution Reaction of Methyl Formate with [Cp′2ZrCl(PHTipp)] – Molecular Structure of meso‐trans ‐[Cp′2ZrCl{OCH(PHTipp)2}] (Cp′ = η5‐C5MeH4, Tipp = 2,4,6‐Pri3C6H2) [Cp′2ZrCl(PHTipp)] ( 1 ) (Cp′ = η5‐C5MeH4, Tipp = 2,4,6‐Pri3C6H2) reacts with methyl formate with insertion and substitution to give [Cp′2ZrCl{OCH(PHTipp)2}] ( 2 ). 2 was characterized spectroscopically (1H, 31P NMR, IR, MS) and by X‐ray structure determination. Only the meso‐trans isomer is present in the solid state.  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