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.
以四丁基碘化铵(BNI) 为有机催化剂, 碘单质(I2) 与偶氮二异庚腈(ABVN) 原位生成的碘代异庚腈为引发剂, 进行甲基丙烯酸甲酯(MMA) 的溶液聚合. 以甲苯为溶剂, MMA:I2:ABVN的摩尔比为200:1:1.7, 考察了催化剂用量对聚合的影响. 结果表明, 加入催化剂可以缩短诱导期, 当I2:BNI摩尔比为1:1时聚合反应的诱导期最短(1.7 h); 当BNI:I2摩尔比为0.25:1~2:1之间时, 聚合物实测分子量与理论值十分接近, 分子量分布较窄, 分子量分布指数(Mw/Mn) 多在1.2以下. 考察了在N,N'-二甲基甲酰胺(DMF)、 四氢呋喃(THF)、 苯甲醚、 苯和甲苯5种溶剂中的聚合反应, 发现在苯和甲苯中聚合可控性最佳, Mw/Mn多在1.2以下; 苯甲醚和THF中聚合速率较快, 聚合物分子量分布较苯中的略宽. 以DMF为溶剂时所得聚合物分子量分布很宽, 聚合可控性差. 核磁共振分析聚合物为碘封端结构, 碘原子封端的聚合物链所占比为91.6%.  相似文献   

2.
The silane-induced ring-opening polymerization of a cyclic siloxane, 1,1,3,3-tetramethyl-2-oxa-1,3-disilacyclopentane (2), is catalyzed by a ruthenium cluster, (μ3235-acenapthylene)Ru3(CO)7 (1), to give poly(tetramethylsilethylenesiloxane) with Mn=6300–780,000 and Mw/Mn=1.5–3.0. The molecular weight of the polymer can be controlled by changing the concentration of the monomer solution. Addition of acetone results in formation of the polymer with Mn=4400, spectroscopic analysis of which reveals existence of a siloxy and an isopropoxy moieties at the end group.  相似文献   

3.
Yb[C(SiMe3)3]2 initiates the living polymerization of methyl methacrylate (MMA) at −78°C to give the polymer with Mn of 51.0×104 (Mw/Mn=1.1) and high isotacticity (97%) in a quantitative yield. Mixing of the acetone solution of resulting polymer (Mn=16.3×104) with the acetone solution of syndiotactic poly(MMA) (Mn=15.7×104) prepared by the (C5Me5)2SmMe(THF) initiator produces desired stereocomplex in high yield bearing very high Tm whose tensile modulus is higher than the respective isotactic and syndiotactic poly(MMA)s. Yb[C(SiMe3)3]2 also generated isotactic (98%) poly[2-(dimethylamino)ethyl methacrylate] (DMEMA), and (C5Me5)2SmMe(THF) affords the syndiotactic (97%) polymer in high yields. The combination of isotactic poly(MMA)-block-poly(DMEMA) (97/3) and syndiotactic poly(MMA)-block-poly(DMEMA) (97/3) provides the amphiphathic stereocomplex. In sharp contrast to the catalysis of Yb[C(SiMe3)3]2 in toluene, the addition of THF or HMPA resulted in the formation of syndio-rich poly(MMA).  相似文献   

4.
With the aim of understanding the influence of donor solvents on the reactivity of the amine complexes [RuCl2(PPh3)2(piperidine)] (1) and [RuCl2(PPh3)2(imidazole)2] (2) in the presence of ethyldiazoacetate, and on the properties of the resulting polymer, a ring opening metathesis polymerization of norbornene was carried out in the presence of small amounts of common solvents such as additives (isopropanol, THF, N,N-dimethylformamide, 2,6-lutidine, isopropanethiol, acetonitrile, dimethyl sulfoxide, NEt3, NH2Me and pyridine). From observations, typical coordinating solvents like DMSO, NEt3, NH2Me and pyridine, hardly affected the yields when either complex was employed. With other additives, the major advantage was the decrease in the polydispersity indices. On using complex 1 with 2,6-lutidine, observed values of Mw/Mn were as low as 1.3, while the yield decreased from 99% to about 20–30% at RT for 1 min in pure solution. In the case of complex 2, which is almost inactive to ROMP (19% at 50 °C for 5 min with Mw/Mn = 6.30), the yield was three-fold (60% at 50 °C for 5 min with Mw/Mn = 1.95) compared to that of without THF. Further, the Mw/Mn was observed to decrease to 1.34 with 200 eq. of THF.  相似文献   

5.
Light scattering measurements in toluene solutions are performed for a series of monodisperse polystyrenes with a molecular weight Mw range from 4×103 to 8×106. The scattered polarized intensities Iv and the natural depolarization ratios ρn are registered with different apparatus at λ=633 or 488 nm and the Mw values are deduced through different formulae. The complete Carr and Zimm formula (CLa), from Iv and ρn, and the usual simplified formula (CLb), from Iv, are considered for the classical method. An already demonstrated formula is considered for the new method (New). Values of Mw and related parameters do not depend on the experimental systems used but deviations appear when using different formulae. The deviations are generally low (about 10%) but often systematic: Mw(CLa)<Mw(CLb)<Mw(New). The most important difference concerns the effect of destructive interferences for Mw>5×105: the new formula leads to a lower increase from θ=90° to θ→0 for Mw values (θ is the observation angle). For instance, in the 8×106 sample, Mw(θ→0)/Mw(θ=90°)=3.6 instead of 6.1, which implies a revision of the usual determination of the radius of gyration, Rg.  相似文献   

6.
[CpR(RPNEt2)]M (CpR=t-BuC5H3, C5(CH3)4, indenyl, fluorenyl; M=Li, K) smoothly react with VCl3(Me3P)2 and CrCl3(THF)3 systems giving paramagnetic complexes [CpR(R1PNEt2)]MCl2 (M=V(Me3P)2, Cr). After reaction with MAO these complexes are active in the polymerisation of ethylene yielding highly crystalline, high-density products of high molecular weight (Mw ranging from 100 000 to 4.5×106 g mol−1, 20≤Tp≤100 °C). Polymerisation with chromium complexes leads to the formation of polyethylenes with broad molecular weight distribution.  相似文献   

7.
使用桥连配体锂盐与MCl4络合, 合成了4个不同结构的双核茂金属化合物[μ,μ-(CH2)3]{[C(H)·(η5-C5H4)(η5-C13H8)](MCl2)}2[M=Zr or Ti](4, 5)和[μ,μ-(CH2)3]{[C(H)(η5-C5H4)(η5-C9H6)]·(MCl2)}2[M=Zr or Ti](6, 7), 配体和化合物都经过核磁氢谱(1H NMR)、 碳谱(13C NMR)、 红外光谱(IR)及元素分析等表征, 确认了化学结构. 以甲基铝氧烷(MAO)为助催化剂, 化合物4~7为催化剂催化丙烯聚合, 考察了聚合温度、 乙烯压力、 铝钛或铝锆比对催化剂活性及聚合物分子量的影响. 结果表明, 多亚甲基桥连双核茂金属是高活性乙烯和丙烯聚合催化剂, 乙烯聚合活性最高达到7.5× 106 g PE/(mol Zr·h)(化合物6), 丙烯聚合活性达 10 × 105 g sPP/(mol Zr·h)(化合物4). 所得间规聚丙烯(sPP)的间规度指数(SI, r) 达到90%. 在同样条件下, 双核化合物的催化活性、 聚合物分子量Mw(> 100000)以及分子量分布(MWD>2.5)均比相应的单核化合物高(Mw<70000, MWD≤2), 表明该体系中存在较强的核效应.  相似文献   

8.
In situ reaction of Li[closo-1-Ph-1,2-C2B10H10] with 7-azabicyclo [4.1.0] heptane results in the formation of the disubstituted carborane, closo-1-Ph-2-(2′-aminocyclohexyl)-1,2-C2B10H10 (1), in 63% yield. Decapitation of (1) with potassium hydroxide in refluxing ethanol produces the cage-opened nido-carborane, K[nido-7-Ph-8-(2′-aminocyclohexyl)-7,8-C2B9H10] (2), in 80% yield. Deprotonation of the above monoanion with two equivalents of n-butyllithium followed by reaction with anhydrous MCl4 · 2THF (M = Zr, Ti) provides d0-half-sandwich metallocarboranes, closo-1-M(Cl)-2-Ph-3-(2′-σ-(H)N-cyclohexyl)-2,3-η5-C2B9H9 (3 M = Zr; 4 M = Ti) in 53% and 42% yields, respectively. The reaction of Li[closo-1,2-C2B10H11] with 7-azabicyclo [4.1.0] heptane in THF affords closo-1-(2′-aminocyclohexyl)-1,2-C2B10H10 (5) in 59% yield. Immobilization of the carboranyl amino ligand (1) to an organic support, Merrifield’s peptide resin (1%), has been achieved by the reaction of the sodium salt of (5) with polystyryl chloride in THF to produce closo-1-(2′-aminocyclohexyl)-2-polystyryl-1,2-C2B10H10 (6) in 87% yield. Further reaction of the dianion derived from (6) with anhydrous ZrCl4 · 2THF led to the formation of the organic polystyryl supported d0-half-sandwich metallocarborane, closo-1-Zr(Cl)-2-(2′-σ-(H)N-cyclohexyl)-3-polystyryl-2,3-η5-C2B9H9 (7), in 38% yield. These new compounds have been characterized by elemental analyses, NMR, and IR spectra. Polymerizations of both ethylene and vinyl chloride with (3) and (7) have been performed in toluene using MMAO-7 (13% ISOPAR-E) as the co-catalyst. Molecular weights up to 32.8 × 103 (Mw/Mn = 1.8) and 9.5 × 103 (Mw/Mn = 2.1) were obtained for PE and PVC, respectively.  相似文献   

9.
Reactions of Co33-CBr)(μ-dppm)(CO)7 with {Au[P(tol)3]}2{μ-(CC)n} (n=2–4) have given {Co3(μ-dppm)(CO)7}{μ33-C(CC)nC} [n=2 (1), 3 (2), 4 (3)] containing carbon chains capped by the cobalt clusters. Tetracyanoethene reacts with 2 to give {Co3(μ-dppm)(CO)7}233-C(CC)2C[=C(CN)2]C[=C(CN)2]C} (4). X-ray structural characterisation of 1, 3 and 4 are reported, that for 3 being the first of a cluster-capped C10 chain.  相似文献   

10.
The structure and texture characteristics of the hybrid organic–inorganic adsorbents, which were obtained by using of two-component systems of “structure-forming agent/trifunctional silane”, are compared as follows: the first component is Si(OC2H5)4 or (C2H5O)3Si–A–Si(OC2H5)3, where A = –(CH2)2– or –C6H4–; the second one is alkoxysilane with amine (–NH2, NH, –NH(CH2)2NH2) and thiol (–SH) groups. The adsorbents, derived from TEOS, have more accessible functional groups (2.6–4.2 mmol/g) than xerogels, which are based on bis(triethoxysilanes) (1.0–2.6 mmol/g). On another hand xerogels derived from bis(triethoxysilanes) have a more extended porous structure (Ssp =516–968 m2/g, Vs = 0.418–1.490 cm3/g, d = 2.5–15.0 nm) than those that are based on TEOS (Ssp = 4–631 m2/g, Vs = 0.005–1.382 cm3/g, d = 2.3–17.7 nm). The geometric dimensions of functional groups have a more essential effect on the parameters of porous structure in the case of TEOS-derived xerogels. Using solid-state NMR spectroscopy, it has been shown that in synthesis of xerogels with the use of TEOS, the molecular frame of globules is formed by structural units Qn (n = 2,3,4), and the functional groups exist as structural units of Tn (n = 2,3). The xerogels obtained with using bis(triethoxysilanes) consist only of structural units of Tn-type (n = 1,2,3).  相似文献   

11.
The ring-opening metathesis polymerization (ROMP) of norbornene catalyzed by bis(acetonitrile) molybdenum and tungsten complexes, [M(η3-C3H5)Cl(CO)2(NCMe)2] (1-Mo: M = Mo, 1-W: M = W), which have two labile acetonitrile ligands, has been investigated. These complexes catalyzed the ROMP of norbornene as a single-component initiator. The highly cis-selective polymerization proceeded in a THF solution (95% for 1-Mo and 96% for 1-W), whereas polymerization in CH2Cl2 or toluene resulted in lower cis selectivity. The polymerization of terminal acetylenes using these complexes was also examined. The tungsten complex 1-W showed a high catalytic activity for the polymerization of terminal acetylenes, such as phenyl- and tert-butylacetylene. A highly active catalytic system for the ROMP of norbornene was achieved by the activation of the tungsten complex, 1-W, with one equivalent of phenylacetylene, giving poly(norbornene) with a high molecular weight (Mn = 391 × 104) and a high cis selectivity (cis  89%).  相似文献   

12.
Novel sul-containing fluorinated polyimides have been synthesized by the reaction of 2,2′-bis-(trifluoromethyl)-4,4′-diaminodiphenyl sulfide (TFDAS) with 1,4-bis-(3,4-dicarboxyphenoxy)benzene dianhydride (HQDPA), 2,2′-bis-(3,4-dicarboxyphenyl)hexafluoropropane dianhydride (6FDA), 4,4′-oxydiphthalicanhydride (ODPA) or 3,4,3′,4′-biphenyl-tetracarboxylic acid dianhydride (s-BPDA). The fluorinated polyimides, prepared by a one-step polycondensation procedure, have good solubility in many solvents, such as N-methyl-2-pyrrolidinone (NMP), dimethylacetamide (DMAc), dimethyl sulfoxide (DMSO), cyclohexanone, tetrahydrofuran (THF) and m-cresol. The molecular weights (Mn's) and polydispersities (Mn/Mw's) of polyimides were in the range of 1.24 × 105 to 3.21 × 105 and 1.59–2.20, respectively. The polymers exhibit excellent thermal stabilities, with glass-transition temperatures (Tg) at 221–275 °C and the 5% weight-loss temperature are above 531 °C. After crosslinking, these polymers show higher thermal stability. The films of polymers have high optical transparency. The novel sul-containing fluorinated polyimides also have low absorption at both 1310 and 1550 nm wavelength windows. Rib-type optical waveguide device was fabricated using the fluorinated polyimides and the near-field mode pattern of the waveguide was demonstrated.  相似文献   

13.
To improve interfacial phenomena of poly(dimethylsiloxane) (PDMS) as biomaterials, well-defined triblock copolymers were prepared as coating materials by reversible addition-fragmentation chain transfer (RAFT) controlled polymerization. Hydroxy-terminated poly(vinylmethylsiloxane-co-dimethylsiloxane) (HO–PVlDmMS–OH) was synthesized by ring-opening polymerization. The copolymerization ratio of vinylmethylsiloxane to dimethylsiloxane was 1/9. The molecular weight of HO–PVlDmMS–OH ranged from (1.43 to 4.44) × 104, and their molecular weight distribution (Mw/Mn) as determined by size-exclusion chromatography equipped with multiangle laser light scattering (SEC-MALS) was 1.16. 4-Cyanopentanoic acid dithiobenzoate was reacted with HO–PVlDmMS–OH to obtain macromolecular chain transfer agents (macro-CTA). 2-Methacryloyloxyethyl phosphorylcholine (MPC) was polymerized with macro-CTAs. The gel-permeation chromatography (GPC) chart of synthesized polymers was a single peak and Mw/Mn was relatively narrow (1.3–1.6). Then the poly(MPC) (PMPC)–PVlDmMS–PMPC triblock copolymers were synthesized. The molecular weight of PMPC in a triblock copolymer was easily controllable by changing the polymerization time or the composition of the macro-CTA to a monomer in the feed. The synthesized block copolymers were slightly soluble in water and extremely soluble in ethanol and 2-propanol.

Surface modification was performed via hydrosilylation. The block copolymer was coated on the PDMS film whose surface was pretreated with poly(hydromethylsiloxane). The surface wettability and lubrication of the PDMS film were effectively improved by immobilization with the block copolymers. In addition, the number of adherent platelets from human platelet-rich plasma (PRP) was dramatically reduced by surface modification. Particularly, the triblock copolymer having a high composition ratio of MPC units to silicone units was effective in improving the surface properties of PDMS.

By selective decomposition of the Si–H bond at the surface of the PDMS substrate by irradiation with UV light, the coating region of the triblock copolymer was easily controlled, resulting in the fabrication of micropatterns. On the surface, albumin adsorption was well manipulated.  相似文献   


14.
Reaction of [Ru3(CO)12 with (CF3)2P---P(CF3)2 in p-xylene at 140°C yielded the compounds [Ru4(CO)13{μ-P(CF3)2}2] (1), [Ru4(CO)14{μ-P(CF3)2}2] (2) and [Ru4(CO)11{μ-P(CF3)2}4] (3). Reaction with [(μ-H)4Ru4(CO)12] under similar conditions yielded [(μ-H)3Ru4(CO)12{μ-P(CF3)2}] (4). All four compounds have been characterised by X-ray crystallography. The fluxional behaviour of the hydrides in 4 has also been studied by variable-temperature NMR spectroscopy. Compounds 1, 2 and 4 were also obtained from the reactions of Ru3(CO)12 with (CF3)2PH in dichloromethane at 80°C.  相似文献   

15.
采用自制的新型双苯并环己酮芳亚胺镍催化剂双苯并环己酮-2,6-二甲基苯亚胺镍(Ⅱ)(Ni{C10H8(O)C[2,6-C6H3(CH3)2N]CH3}2, C1)和双苯并环己酮-2,6-二氯苯亚胺镍(Ⅱ)(Ni{C10H8(O)C[2,6-C6H3Cl2N]CH3}2, C2)与三五氟苯硼[B(C6F5)3]结合, 在一定的反应条件下可高效催化降冰片烯(NB)与甲基丙烯酸正丁酯(n-BMA)的乙烯基加成共聚合. 提出了催化聚合时存在的可能失活机理; 研究了不同单体投料比对催化活性、 产率及产物性能的影响. 根据Kelen-Tüdõs方法分别估算出2种单体在不同催化体系下的竞聚率, 即当催化体系为C1/B(C6F5)3时, 竞聚率rn-BMA=0.02, rNB=16.28, rNB·rn-BMA=0.32; 当催化体系为C2/B(C6F5)3时, rn-BMA=0.01, rNB=64.83, rNB·rn-BMA=0.65. 结果表明, 2种单体在2种体系催化下均为无规共聚合.  相似文献   

16.
This work presents chemical modeling of solubilities of metal sulfates in aqueous solutions of sulfuric acid at high temperatures. Calculations were compared with experimental solubility measurements of hematite (Fe2O3) in aqueous ternary and quaternary systems of H2SO4, MgSO4 and Al2(SO4)3 at high temperatures. A hybrid model of ion-association and electrolyte non-random two liquid (ENRTL) theory was employed to fit solubility data in three ternary systems H2SO4–MgSO4–H2O, H2SO4–Al2(SO4)3–H2O at 235–270 °C and H2SO4–Fe2(SO4)3–H2O at 150–270 °C. Employing the Aspen Plus™ property program, the electrolyte NRTL local composition model was used for calculating activity coefficients of the ions Al3+, Mg2+ Fe3+ and SO42−, HSO4, OH, H3O+, respectively, as well as molecular species. The solid phases were hydronium alunite (H3O)Al3(SO4)2(OH)6, hematite Fe2O3 and magnesium sulfate monohydrate (MgSO4)·H2O which were employed as constraint precipitation solids in calculating the metal sulfate solubilities. A correlation for the equilibrium constants of the association reactions of complex species versus temperature was implemented. Based on the maximum-likelihood principle, the binary interaction energy parameters for the ionic species as well as the coefficients for equilibrium constants of the reactions were obtained simultaneously using the solubility data of the ternary systems. Following that, the solubilities of metal sulfates in the quaternary systems H2SO4–Fe2(SO4)3–MgSO4–H2O, H2SO4–Fe2(SO4)3–Al2(SO4)3–H2O at 250 °C and H2SO4–Al2(SO4)3–MgSO4–H2O at 230–270 °C were predicted. The calculated results were in excellent agreement with the experimental data.  相似文献   

17.
[1,8-C10H6(NR)2]TiCl2 (3; R=SiMe3, SiiBuMe2, SiiPr3) complexes have been prepared from dilithio salts [1,8-C10H6(NR)2]Li2 (2) and TiCl4 in diethyl ether in moderate yields (60–63%). These complexes showed significant catalytic activities for ethylene polymerization and for ethylene/1-hexene copolymerization in the presence of methylaluminoxane (MAO), methyl isobutyl aluminoxane (MMAO), AliBu3– or AlEt3–Ph3CB(C6F5)4 as a cocatalyst. The catalytic activities performed in heptane (cocatalyst MMAO) were higher than those carried out in toluene (cocatalyst MAO): 709 kg-PE/mol-Ti·h could be attained for ethylene polymerization by using [1,8-C10H6(NSiiBuMe2)2]TiCl2–MMAO catalyst system.  相似文献   

18.
The reaction of 1,2-bis(diphenylthioylphosphino)hydrazine (L) with copper(I) and mercury(II) halides affords the complexes, [{CuLX}2] (X = I, Br or Cl), [HgLX2] (X = Cl or Br) and the tetrametallic complex, [{L(HgI2)2}2]. Single crystal X-ray structures have been performed on the uncoordinated ligand, L, as well as the complexes [{CuLX}2] (X = I, Br and Cl), [HgLBr2] and [{L(HgI2)2}2. The molecules of L exist as dimers as a result of pairs of N–HS hydrogen bonds. The copper(I) complexes are centrosymmetric dimetallic species, the two copper atoms being bridged by L and the X atoms. In all cases the coordination sphere around the Cu atoms is approximately trigonal pyramidal with an ‘S2X2’ donor set. The complex, [HgLBr2], is a distorted tetrahedral monomer with an ‘S2Br2’ donor set and L acting as a bidentate thus forming a seven-membered chelate ring. The tetramercury iodo complex, [{L(HgI2)2}2], contains two ‘L(HgI2)2’ units linked centrosymmetrically via an I atom from each moiety. The geometry around the Hg atoms is distorted tetrahedral. The influence of hydrogen bonding between the hydrazine backbone hydrogens of L and the coordinated halide ions in for the structures of the metal complexes is discussed.  相似文献   

19.
139La-NMR chemical shifts were measured for several anionic complexes of formulae Li(C4H8O2)3/2 [La(ν3-C3H5)4], [Li(C4H8O2)2][Cp′nLa(ν3-C3]H5)4−n] (Cp′ = Cp(ν5-C5H5); n = 1, 2 and Cp′ = Cp * (ν5-C5Me5); N = 1) and Li[RnLa(ν3-C3H4)4n] (R = N(SiMe3)2; n = 1, 2 and R = CCsIMe3; n = 4), as well as for neutral compounds for formulae La(ν3-C3H5)3Ln (L = (C4H8O2)1.5, (HMPT)2, TMED), Cp′nLa(ν3-C3H5)3−n (Cp′= Cp(ν5-Cp5H5), Cp *(ν5-C5Me5); n = 1, 2) and La(ν3-C3H2)2X(THF)2 X = Cl, Br, I). Typical ranges of the 139La-NMR chemical shifts were found for the different types of complex independent of number and kind of organyl groups directly bonded to lanthanum.

Zusammenfassung

139La-NMR-Spektroskopie wurde an einer Reihe anionischer Allyllanthanat(III)-Komplexe der Zusammensetzung ]- [La)ν3-C3H5)4, [Li(C4H8)2][Cp′nLa(ν3-C3H5)4−n(Cp′ = Cp(ν5-C5H5); n = 1, 2 und Cp′ = Cp * (ν5-C5Me5); N = 1) und Li[RnLa(ν3-C3H5)4−n (R = B(SiMe3)2; n = 1, 2 und R = CCSiMe3; n = 4 sowie neutraler Allyllanthan(III)-Komplexe der Zusammensetzung La(ν3-C3H5)3Ln (Ln = (C4H8O2)1.5, (HMPT)2, TMED), Cp′n, La(ν3-C3H5)3−n (Cp′ = Cp(ν5-C5H5), Cp * (ν5- Cp5Me5); n = 1, 2) und La(ν3-Cp3H5)2X(THF)2 (X = Cl, Br, I) durchgefürt. In Abhängikeit von der Anzahl und der Art der am Lanthan gebundenen Gruppen wurden für die verschieden Komplextypen charakteristische Resonanzbereiche ermittelt.  相似文献   


20.
The interaction of [(η5-C5H4But)2YbCl · LiCl] with one equivalent of Li[(CH2) (CH2)PPh2] in tetrahydrofuran gave [Ph2PMe2][(η5-C5H4But)2Li] (1) and [(η5-C5H4But)2Yb(Cl)CH2P(Me)Ph2] (2) in 10% and 30% yields, respectively. 1 could also be prepared in 70% yield from the reaction of [Ph2PMe2][CF3SO3] with two equivalents of (C5H4But)Li. Both compounds have been fully characterized by analytical, spectroscopic and X-ray diffraction methods. The solid state structure of 1 reveals a sandwich structure for the [(η5-C5H4But)2Li] anion.  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