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 581 毫秒
1.
Drastic effects of Lewis acids E(C6F5)3 (E = Al, B) on polymerization of functionalized alkenes such as methyl methacrylate (MMA) and N,N-dimethyl acrylamide (DMAA) mediated by metallocene and lithium ester enolates, Cp2Zr[OC(OiPr)CMe2]2 (1) and Me2CC(OiPr)OLi, are documented as well as elucidated. In the case of metallocene bis(ester enolate) 1, when combined with 2 equiv. of Al(C6F5)3, it effects highly active ion-pairing polymerization of MMA and DMAA; the living nature of this polymerization system allows for the synthesis of well-defined diblock and triblock copolymers of MMA with longer-chain alkyl methacrylates. In sharp contrast, the 1/2B(C6F5)3 combination exhibits low to negligible polymerization activity due to the formation of ineffective adduct Cp2Zr[OC(OiPr)CMe2]+[OC(OiPr)CMe2B(C6F5)3] (2). Such a profound Al vs. B Lewis acid effect has also been observed for the lithium ester enolate; while the Me2CC(OiPr)OLi/2Al(C6F5)3 system is highly active for MMA polymerization, the seemingly analogous Me2CC(OiPr)OLi/2B(C6F5)3 system is inactive. Structure analyses of the resulting lithium enolaluminate and enolborate adducts, Li+[Me2CC(OiPr)OAl(C6F5)3] (3) and Li+[Me2CC(OiPr)OB(C6F5)3] (4), coupled with polymerization studies, show that the remarkable differences observed for Al vs. B are due to the inability of the lithium enolborate/borane pair to effect the bimolecular, activated-monomer anionic polymerization as does the lithium enolaluminate/alane pair.  相似文献   

2.
The dimesitylpropargylphosphanes mes2P?CH2?C≡C?R 6 a (R=H), 6 b (R=CH3), 6 c (R=SiMe3) and the allene mes2P?C(CH3)=C=CH2 ( 8 ) were reacted with Piers’ borane, HB(C6F5)2. Compound 6 a gave mes2PCH2CH=CH(B(C6F5)2] ( 9 a ). In contrast, addition of HB(C6F5)2 to 6 b and 6 c gave mixtures of 9 b (R=CH3) and 9 c (R=SiMe3) with the regioisomers mes2P?CH2?C[B(C6F5)2]=CRH 2 b (R=CH3) and 2 c (R=SiMe3), respectively. Compounds 2 b , c underwent rapid phosphane/borane (P/B) frustrated Lewis pair (FLP) reactions under mild conditions. Compound 2 c reacted with nitric oxide (NO) to give the persistent FLP NO radical 11 . The systems 2 b , c cleaved dihydrogen at room temperature to give the respective phosphonium/hydridoborate products 13 b , c . Compound 13 c transferred the H+/H? pair to a small series of enamines. Compound 13 c was also a metal‐free catalyst (5 mol %) for the hydrogenation of the enamines. The allene 8 reacted with B(C6F5)3 to give the zwitterionic phosphonium/borate 17 . The ‐PPh2‐substituted mes2P‐propargyl system 6 d underwent a typical 1,2‐P/B‐addition reaction to the C≡C triple bond to form the phosphetium/borate zwitterion 20 . Several products were characterized by X‐ray diffraction.  相似文献   

3.
The preparation of a new functionalized cyclopentadienyl ligand bearing a nitrile pendant substituent, (C5H4CMe2CH2CN)? is reported. The corresponding lithium salt of this ligand (1) was prepared by the reaction of in situ lithiated acetonitrile with 6,6-dimethylfulvene. The ligand was subsequently utilized for the synthesis of group 4 metal complexes [(η5–C5H4CMe2CH2CN)2MCl2] (M = Ti, 2; M = Zr, 3; M = Hf, 4), [(η5–C5H5) (η5–C5H4CMe2CH2CN)MCl2] (M = Ti, 7; M = Zr, 8), and [(η5-C5Me5) (η5 C5H4CMe2CH2CN)2ZrCl2] (9). Alternative route to 2 comprised the preparation of half-sandwich complex [(η5–C5H4CMe2CH2CN)TiCl3] (6). The prepared compounds were characterized by common spectroscopic methods and the solid state structures of complexes 2, 3, 4, 7, and 9 were determined by the single-crystal X-ray diffraction analysis. In addition, compound 7 was converted to the corresponding dimethyl derivative [(η5–C5H5) (η5–C5H4CMe2CH2CN)TiMe2] (10) and also treated with the chloride anion abstractor Li[B(C6F5)4] to generate the cationic complex with the coordinated nitrile group, as suggested by the NMR spectroscopy. A formation of yet another cationic complex was observed upon treating compound 10 with (Ph3C)[B(C6F5)4].  相似文献   

4.
Imines, Im, such as MeN=C(Ph)H (5), 2-methyl 4,5-dihydrothiazole (8a), 2-methyl 4,5-dihydrooxazole (8b) and MeN=C(OMe)Me (13) add to the α-carbon atom of the vinylidene ligand in [(CO)5Cr=C=CMe2] (4) to give isolable zwitterionic adducts, [(CO)5Cr–C(=CMe2)(Im+)]. The reaction of [(CO)5W=C=CPh2] (12) with 13 also yields an adduct, [(CO)5W–C(=CPh2){NMe=C(OMe)Me}+] (15), whereas from the corresponding reaction of 4 with xanthylideneimine, H–N=C(C6H4)2O (16), a carbene complex, [(CO)5Cr=C(i-Pr)–N=C(C6H4)2O] (17), is obtained. Complex 17 presumably is formed by initial addition of 16 to 4 and subsequently rapid rearrangement. In solution, the adduct [(CO)5Cr–C(=CMe2)(NMe=C(Ph)H)+] (6) slowly cyclizes to form the 2-azetidin-1-ylidene complex [(CO)5Cr= Me2] (7). In contrast, when solution of those zwitterions are heated that are formed by addition of 4,5-dihydrothiazole or 4,5-dihydrooxazole to 4, no cyclization is observed but rather the formation of 4,5-dihydrothiazole and 4,5-dihydrooxazole complexes, respectively. The structures of two adducts, [(CO)5Cr–C(=CMe2)(Im+)] (Im=MeN=C(Ph)H, 2-methyl 4,5-dihydrothiazole) and of the substitution product [(CO)5W(2-methyl 4,5-dihydrothiazole)] have been established by X-ray structural analyses.  相似文献   

5.
(C6F5)3Sb has been found to react with interhalogens and halo-pseudohalogens, IX(X = Cl, Br, N3 and NCO), pseudohalogen (SCN), and elemental sulphur to give oxidative addition products (I–VI). (C6F5)3SbS(VI) may also be prepared by the reaction of (C6F5)3SbCl2 with H2S. Metathetical reactions of (C6F5)3SbCl2 with appropriate metallic salts yield covalent pentacoordinate disubstituted products (V, VII–XII) of the general formula, (C6F5)3SbY2 (Y = NCS, NCO, ?ONCMe2, ?ONCMePh ?NCO(CH2)2CO and p-NO2C6H4OCO). Treatment of (C6F5)3SbCl2 with aqueous NaN3 gives the binuclear oxo-bridge compound, [(C6F5)3SbOSb(C6F5)3](N3)2·(III) and (IV) are also accessible by displacement reaction of (I) or (II) with the corresponding metallic salt. Molecular weight, conductance measurements, and IR spectra on the new organoantimony(V) derivatives have been obtained.Reductive cleavage reactions of (C6F5)3SbS with hexaaryldileads, Ar6Pb2(Ar = Phenyl, p-tolyl) produce (C6F5)3Sb and the corresponding bis(triaryllead) sulphide but treatment of (C6F5)3SbX2(X = NCO, Cl) with Ar6Pb2 gave Ar4Pb and Ar2PbX2 together with (C6F5)3Sb.(C6F5)3SbCl2 and bis(triorganotin)sulphides undergo exchange of anionic groups.  相似文献   

6.
Herein, we present the formation of transient radical ion pairs (RIPs) by single-electron transfer (SET) in phosphine−quinone systems and explore their potential for the activation of C−H bonds. PMes3 (Mes=2,4,6-Me3C6H2) reacts with DDQ (2,3-dichloro-5,6-dicyano-1,4-benzoquinone) with formation of the P−O bonded zwitterionic adduct Mes3P−DDQ ( 1 ), while the reaction with the sterically more crowded PTip3 (Tip=2,4,6-iPr3C6H2) afforded C−H bond activation product Tip2P(H)(2-[CMe2(DDQ)]-4,6-iPr2-C6H2) ( 2 ). UV/Vis and EPR spectroscopic studies showed that the latter reaction proceeds via initial SET, forming RIP [PTip3]⋅+[DDQ]⋅, and subsequent homolytic C−H bond activation, which was supported by DFT calculations. The isolation of analogous products, Tip2P(H)(2-[CMe2{TCQ−B(C6F5)3}]-4,6-iPr2-C6H2) ( 4 , TCQ=tetrachloro-1,4-benzoquinone) and Tip2P(H)(2-[CMe2{oQtBu−B(C6F5)3}]-4,6-iPr2-C6H2) ( 8 , oQtBu=3,5-di-tert-butyl-1,2-benzoquinone), from reactions of PTip3 with Lewis-acid activated quinones, TCQ−B(C6F5)3 and oQtBu−B(C6F5)3, respectively, further supports the proposed radical mechanism. As such, this study presents key mechanistic insights into the homolytic C−H bond activation by the synergistic action of radical ion pairs.  相似文献   

7.
Protocols for the synthesis of the bulky polyfluorinated triarylboranes 2,6-(C6F5)2C6F3B(C6F5)2 ( 1 ), 2,6-(C6F5)2C6F3B[3,5-(CF3)2C6H3] ( 2 ), 2,4,6-(C6F5)3C6H2B(C6F5)2 ( 3 ), 2,4,6-(C6F5)3C6H2B[3,5-(CF3)2C6H3] ( 4 ) were developed. All boranes are water tolerant and according to the Gutmann-Beckett method, 1 – 3 display Lewis acidities larger than that of the prominent B(C6F5)3.  相似文献   

8.
O-Halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines – Synthesis, Crystal Structure, and Reactions The substitution of halogenosilanes on lithiated N,O-bis(trimethylsilyl)-hydroxylamine in the molar ratio of 1 : 1 occurs on the oxygen atom. The O-halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines were prepared: RSiF2ON · (SiMe3)2 (R = CMe3 1 , CHMe2 2 , CH2C6H5 3 , C6H2(CMe3)3 4 ), RR′SiFON(SiMe3)2 (R = CMe3, R′ = C6H5 5 ; R = Me, R′ = C6H5 6 ; R = C6H2Me3, R′ = C6H2Me3 7 ; R = CH2C6H5, R′ = CH2C6H5 8 ; R = CHMe2, R′ = CHMe2 9 ; R = CMe3, R′ = CMe3 10 ), RSiCl2ON(SiMe3)2 (R = CMe3 11 ; R = Cl 12 ). The reaction of fluorosilanes with lithiated N,O-bis(trimethylsilyl)hydroxylamine in the molar ratio of 1 : 2 leads to the formation of O,O′-fluorosilyl-bis[N,N-bis(trimethylsilyl)hydroxylamines]: RSiF[ON(SiMe3)2]2 (R = CMe3 13 ; R = C6H5 14 ). 13 could be prepared in the reaction of 1 with LiON(SiMe3)2. Lithiated dimethylketonoxime reacts with 1 to Me2C=NOSiRF–ON(SiMe3)2 [R = CMe3 ( 15 )]. The first crystal structure of a tris(silyl)hydroxylamine ( 4 ) is shown. The angle at the nitrogen prove a pyramidal geometry.  相似文献   

9.
The perfluoroaryl tellurolates C6F5TeLi (1) and 4-CF3C6F4TeLi (2) were prepared. These intermediates were identified by NMR spectroscopy and may form, depending on the reaction conditions, either the corresponding ditellanes C6F5TeTeC6F5 (3) and CF3C6F4TeTeC6F4CF3 (4) by subsequent oxidation, or in the case of 1, a telluranthrene (C6F4Te)2 (5) by reaction with itself. The halogenation products of 5, ( C6F4Te)2F4 (6), (C6F4Te)2Cl4 (7), (C6F4Te)2Br4 (8), as well as the azidation product (C6F4Te)2(N3)4 (9) were synthesized. Furthermore, in pursuit of our recent work on tellurium azides, the syntheses and properties of R2Te(N3)2 (R=CF3 (10), C6F2H3 (11)) and RTe(N3)3 (R=CF3 (12) and C6F5 (13)) are reported. The crystal structures of CF3C6F4TeTeC6F4CF3 (4), (C6F4Te)2Br4 (8), and (C6F2H3)2Te(N3)2 (11) were determined.  相似文献   

10.
Phosphorus‐bridged strained [1]ferrocenophanes [Fe{(η‐C5H4)2P(CH2CMe3)}] ( 2 ) and [Fe{(η‐C5H4)2P(CH2SiMe3)}] ( 3 ) with neopentyl and (trimethylsilyl)methyl substituents on phosphorus, respectively, have been synthesized and characterized. Photocontrolled living anionic ring‐opening polymerization (ROP) of the known phosphorus‐bridged [1]ferrocenophane [Fe{(η‐C5H4)2P(CMe3)}] ( 1 ) and the new monomers 2 and 3 , initiated by Na[C5H5] in THF at 5 °C, yielded well‐defined polyferrocenylphosphines (PFPs), [Fe{(η‐C5H4)2PR}]n (R=CMe3 ( 4 ), CH2CMe3 ( 5 ), and CH2SiMe3 ( 6 )), with controlled molecular weights (up to ca. 60×103 Da) and narrow molecular weight distributions. The PFPs 4 – 6 were characterized by multinuclear NMR spectroscopy, DSC, and by GPC analysis of the corresponding poly(ferrocenylphosphine sulfides) obtained by sulfurization of the phosphorus(III) centers. The living nature of the photocontrolled anionic ROP allowed the synthesis of well‐defined all‐organometallic PFP‐b‐PFSF ( 7 a and 7 b ) (PFSF=polyferrocenylmethyl(3,3,3,‐trifluoropropyl)silane) diblock copolymers through sequential monomer addition. TEM studies of the thin films of the diblock copolymer 7 b showed microphase separation to form cylindrical PFSF domains in a PFP matrix.  相似文献   

11.
[(BDI)Mg+][B(C6F5)4] ( 1 ; BDI=CH[C(CH3)NDipp]2; Dipp=2,6-diisopropylphenyl) was prepared by reaction of (BDI)MgnPr with [Ph3C+][B(C6F5)4]. Addition of 3-hexyne gave [(BDI)Mg+ ⋅ (EtC≡CEt)][B(C6F5)4]. Single-crystal X-ray analysis, NMR investigations, Raman spectra, and DFT calculations indicate a significant Mg-alkyne interaction. Addition of the terminal alkynes PhC≡CH or Me3SiC≡CH led to alkyne deprotonation by the BDI ligand to give [(BDI-H)Mg+(C≡CPh)]2 ⋅ 2 [B(C6F5)4] ( 2 , 70 %) and [(BDI-H)Mg+(C≡CSiMe3)]2 ⋅ 2 [B(C6F5)4] ( 3 , 63 %). Addition of internal alkynes PhC≡CPh or PhC≡CMe led to [4+2] cycloadditions with the BDI ligand to give {Mg+C(Ph)=C(Ph)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 4 , 53 %) and {Mg+C(Ph)=C(Me)C[C(Me)=NDipp]2}2 ⋅ 2 [B(C6F5)4] ( 5 , 73 %), in which the Mg center is N,N,C-chelated. The (BDI)Mg+ cation can be viewed as an intramolecular frustrated Lewis pair (FLP) with a Lewis acidic site (Mg) and a Lewis (or Brønsted) basic site (BDI). Reaction of [(BDI)Mg+][B(C6F5)4] ( 1 ) with a range of phosphines varying in bulk and donor strength generated [(BDI)Mg+ ⋅ PPh3][B(C6F5)4] ( 6 ), [(BDI)Mg+ ⋅ PCy3][B(C6F5)4] ( 7 ), and [(BDI)Mg+ ⋅ PtBu3][B(C6F5)4] ( 8 ). The bulkier phosphine PMes3 (Mes=mesityl) did not show any interaction. Combinations of [(BDI)Mg+][B(C6F5)4] and phosphines did not result in addition to the triple bond in 3-hexyne, but during the screening process it was discovered that the cationic magnesium complex catalyzes the hydrophosphination of PhC≡CH with HPPh2, for which an FLP-type mechanism is tentatively proposed.  相似文献   

12.
Synthesis of Mono- and Bis(silyl)hydroxylamines Silylamines reacts with hydroxylaminehydrochlorid to give the monosilylhydroxylamines: R2FSiONH2 (R = CMe3 1 ), R2R′SiONH2 (R = CMe3, R′ = Me 2 ), R2(NH2)SiONH2 (R = CMe3 3 ). The reaction of 1 in the present of HCl-acceptors or the reaction of lithiated 1 with Me3SiCl or F2Si(CMe3)2 leads to the formation of bis(silyl)hydroxylamines, (Me3C)2FSiONHSiMe3 4 , and (Me3C)2FSiONHSiF(CMe3)2 5 . The lithium derivatives of Me3SiONH2 and 2 react with fluorosilanes to the bis(silyl)hydroxylamines: Me3SiONHSiFRR′ (R = R′ = CMe3, 6 , R = CMe3, R′ = F 7 , R = R′ = NMeSiMe3 8 ), (Me3C)2MeSiNHOSiFRR′ (R = CMe3, R′ = F 9 , R = (Me3C)3C6H2, R′ = F 10 , R = R′ = CMe3 11 , R = R′ = CHMe2 12 ). The bis(silyl)hydroxylamines 4 and 6 are structure isomers.  相似文献   

13.
Tris(pentafluorophenyl)silylamines were synthesized by silylation of amines and imines with (C6F5)3SiCl or (C6F5)3SiOTf in the presence of triethylamine. The crystal structures of the (C6F5)3SiN(H)CH2Ph and (C6F5)3SiN(CH=CMe2)CH2Ph compounds were studied by X-ray diffraction. The crystal packings were analyzed by quantum chemical calculations in terms of the density functional theory (PBE exchange-correlation functional). Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1345–1352, July, 2007.  相似文献   

14.
The reaction of (C6F5)2BH ( 1 ) with N,N‐dimethylallylamine ( 2 ), N,N‐diethylallylamine ( 3 ) and 1‐allylpiperidine ( 4 ) afforded the five‐membered ring systems (C6F5)2B(CH2)3NR2 (R=Me ( 5 ), Et ( 6 )) and (C6F5)2B(CH2)3N(CH2)5 ( 7 ) with an intramolecular dative B? N bond. A different product was obtained from the reaction of (C6F5)2BH ( 1 ) with N,N‐diisopropylallylamine ( 8 ), which afforded the seven‐membered ring system (C6F5)2B(CH2)3N(iPr)CH(Me)CH2 ( 9 ) under extrusion of dihydrogen. All compounds were characterised by elemental analysis, NMR spectroscopy and single‐crystal X‐ray diffraction experiments. Density functional theory (DFT) studies were performed to rationalise the different reaction mechanism for the formation of products 6 and 9 . The bonding situation of compound 9 was analysed in terms of its electron density topology to describe the delocalised nature of a borane– enamine adduct.  相似文献   

15.
Molybdenum(VI) bis(imido) complexes [Mo(NtBu)2(LR)2] (R=H 1 a ; R=CF3 1 b ) combined with B(C6F5)3 ( 1 a /B(C6F5)3, 1 b /B(C6F5)3) exhibit a frustrated Lewis pair (FLP) character that can heterolytically split H−H, Si−H and O−H bonds. Cleavage of H2 and Et3SiH affords ion pairs [Mo(NtBu)(NHtBu)(LR)2][HB(C6F5)3] (R=H 2 a ; R=CF3 2 b ) composed of a Mo(VI) amido imido cation and a hydridoborate anion, while reaction with H2O leads to [Mo(NtBu)(NHtBu)(LR)2][(HO)B(C6F5)3] (R=H 3 a ; R=CF3 3 b ). Ion pairs 2 a and 2 b are catalysts for the hydrosilylation of aldehydes with triethylsilane, with 2 b being more active than 2 a . Mechanistic elucidation revealed insertion of the aldehyde into the B−H bond of [HB(C6F5)3]. We were able to isolate and fully characterize, including by single-crystal X-ray diffraction analysis, the inserted products Mo(NtBu)(NHtBu)(LR)2][{PhCH2O}B(C6F5)3] (R=H 4 a ; R=CF3 4 b ). Catalysis occurs at [HB(C6F5)3] while [Mo(NtBu)(NHtBu)(LR)2]+ (R=H or CF3) act as the cationic counterions. However, the striking difference in reactivity gives ample evidence that molybdenum cations behave as weakly coordinating cations (WCC).  相似文献   

16.
B(C6F5)3 and P(MeNCH2CH2)3N form a classical Lewis adduct, (C6F5)3BP(MeNCH2CH2)3N. Although (C6F5)3BP(MeNCH2CH2)3N does not exhibit spectroscopic evidence of dissociation into its constituent acid and base, products of frustrated Lewis pair (FLP) addition reactions are seen with PhNCO, PhCH2N3, PhNSO, and CO2. Computational studies show that thermal access to the dissociated acid and base permits FLP reactivity to proceed. These results demonstrate that FLP reactivity extends across the entire continuum of equilibria governing Lewis acid‐base adducts.  相似文献   

17.
The frustrated Lewis pair (FLP) Mes2PCH2CH2B(C6F5)2 ( 1 ) reacts with an enolizable conjugated ynone by 1,4‐addition involving enolate tautomerization to give an eight‐membered zwitterionic heterocycle. The conjugated endione PhCO‐CH?CH‐COPh reacts with the intermolecular FLP tBu3P/B(C6F5)3 by a simple 1,4‐addition to an enone subunit. The same substrate undergoes a more complex reaction with the FLP 1 that involves internal acetal formation to give a heterobicyclic zwitterionic product. FLP 1 reacts with dimethyl maleate by selective overall addition to the C?C double bond to give a six‐membered heterocycle. It adds analogously to the triple bond of an acetylenic ester to give a similarly structured six‐membered heterocycle. The intermolecular FLP P(o‐tolyl)3/B(C6F5)3 reacts analogously with acetylenic ester by trans‐addition to the carbon–carbon triple bond. An excess of the intermolecular FLP tBu3P/B(C6F5)3, which contains a more nucleophilic phosphane, reacts differently with acetylenic ester examples, namely by O? C(alkyl) bond cleavage to give the {R‐CO2[B(C6F5)3]2?}[alkyl‐PtBu3+] salts. Simple aryl or alkyl esters react analogously by using the borane‐stabilized carboxylates as good leaving groups. All essential products were characterized by X‐ray diffraction.  相似文献   

18.
1,2,3,4,7,7-Hexafluorobicyclo[2.2.1]heptadiene (1) and 2,3-bis(trimethyltin)-1,4,5,6,7,7-hexafluorobicyclo[2.2.1]hepta-2,5-diene (2) react with [M(Ph3P)4] (M = Pt, Pd) to afford air-stable adducts. 2,3-Dichloro-1,4,5,6,7,7-hexafluorobicyclo[2.2.1]hepta-2,5-diene (3) gives only [PtCl2(PPh3)2] with [Pt(Ph3P)4], but a low yield of an adduct was obtained with [Pd(PPh3)4]. The diene 1 also reacts with Fe(CO)5 to form the complex [(C7H2F6)Fe(CO)4], and with [Rh(C2H4)2(acac)] to give [(C7H2F6)Rh(acac)] in which the diene acts as a bidentate ligand. Similar products could not be isolated from the reactions of 2 and 3. A stable adduct, believed to be [{C7F6(SnMe3)2}Rh(CO)2(μ-Cl)2Rh(CO)2] has been isolated from the reaction between 2 and [Rh(CO)2Cl]2. This adduct reacts with PPh3 to give the bridge-cleavage product [{C7F6(SnMe3)2}RhCl(CO)(PPh3)2]. Reaction of 1 with [Rh(CO)2Cl]2 gives an unstable adduct which could not be isolated, and 2 does not react at room temperature. The chloro derivative 3 reacts with [PdCl2(PhCN)2] to give the adduct [(C7F6Cl2)PdCl(PhCN)], but 1 and 2 do not react under similar conditions. Stable substitution products [(C7F6R2)M] (R = H, M = Fe(CO)2(η-C5H5); R = SnMe3, M = Fe(CO)2(η-C5H5), Mn(CO)5, Ir(CO)2(PPh3)2, Rh(CO)2(PPh3)2; R = Cl, M = Ir(CO)2(PPh3)2, Rh(CO)2(PPh3)2) have been isolated from the reactions of the dienes with carbonylmetal anions. Insertion of the CHCH bond occurs when 1 is heated with [MnMe(CO)5] to give [{C7F6H2C(O)Me}Mn(CO)4], and this, on reaction with either PPh3 or [Pt(PPh3)4], gives [(C7F6H2COMe)Mn(CO)4PPh3].  相似文献   

19.
《Tetrahedron: Asymmetry》2003,14(7):787-790
A series of racemic dirhodium(II) compounds with two ortho-metalated aryl phosphine ligands in a head-to-tail arrangement Rh2(O2CR)2(pc)2 (pc=ortho-metalated aryl phosphine) (1ak) were tested in the regio- and stereoselective cyclopropanation of racemic 1-diazo-6-methyl-3-(2-propenyl)-5-hepten-2-one 2, which possesses two different reactive CC double bonds for a five-membered-ring formation. The complexes Rh2(O2CCH3)2(pc)2 {pc=[(C6H4)P(C6H5)2], [(p-CH3C6H3)P(p-CH3C6H4)2], and [(C6H4)P(C6H5)(C6F5)]} (1ad) successfully enhanced the cyclopropanation of trisubstituted versus monosubstituted CC bonds to give an 80:20 selectivity ratio. The reaction occurred with excellent diastereoselectivity; the syn-products were the only stereoisomers observed in the whole series of the catalysts. Enantioenriched products were obtained when enantiomerically pure dirhodium(II) complexes were used.  相似文献   

20.
S-Nitrosothiols (RSNOs) serve as air-stable reservoirs for nitric oxide in biology. While copper enzymes promote NO release from RSNOs by serving as Lewis acids for intramolecular electron-transfer, redox-innocent Lewis acids separate these two functions to reveal the effect of coordination on structure and reactivity. The synthetic Lewis acid B(C6F5)3 coordinates to the RSNO oxygen atom, leading to profound changes in the RSNO electronic structure and reactivity. Although RSNOs possess relatively negative reduction potentials, B(C6F5)3 coordination increases their reduction potential by over 1 V into the physiologically accessible +0.1 V vs. NHE. Outer-sphere chemical reduction gives the Lewis acid stabilized hyponitrite dianion trans-[LA-O-N=N-O-LA]2− [LA=B(C6F5)3], which releases N2O upon acidification. Mechanistic and computational studies support initial reduction to the [RSNO-B(C6F5)3] radical anion, which is susceptible to N−N coupling prior to loss of RSSR.  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