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.
The initiation stage of methyl methacrylate polymerization in the presence of benzoyl peroxide-metallocene (MC) systems is considered, with MC = Cp2Fe, (C5Me5)2Fe, (AcC5H4)(C5H5)Fe, Cp2TiCl2, Cp2ZrCl2, and (C5Me5)2ZrCl2. The decomposition of benzoyl peroxide in the presence of a metallocene proceeds via the formation of its complex with the metallocene. The catalytic effect of the metallocenes on the initiation of methyl methacrylate polymerization is due to the formation of a metallocene-benzoyl peroxide complex and its decomposition yielding primary radicals. The chain propagation stage is metallocene-dependent, which is explained by the formation of complex sites. Their formation pathway and structures are analyzed using quantum-chemical calculations.  相似文献   

2.
Methyl methacrylate was polymerized with Cp2YCl(THF) or IVB group metallocene compounds (i.e., Cp2ZrCl2 and Cp2HfCl2, etc.), in the presence of a Lewis acid like Zn(C2H5)2. The Lewis acid was complexed with methyl methacrylate, which avoided the metallocene compounds being poisoned with a functional group. A living polymerization was promoted through the use of metallocene/MAO/Zn(C2H5)2, which gave tactic poly(methyl methacrylate) with a high molecular weight. The polymer yield increases with polymerization time, which indicates that the propagation rate is zero in order in the concentration of the monomer. The polymer yield increases also with the concentration of Cp2YCl(THF), which indicates the yttrocene to be the real catalyst. When the polymerization temperature exceeds room temperature, the poly(methyl methacrylate) cannot be synthesized by the Cp2YCl(THF) catalyst. When the reaction temperature reachs −60 °C, the poly(methyl methacrylate) is high syndiotatic and molecular weight by the Cp2YCl(THF)/MAO catalyst system. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 1184–1194, 2000  相似文献   

3.
The influence of differently substituted cyclopentadienyl CpR ligands on the reaction outcome of [CpRFe(CO)2]2 (CpR = C5Me5, EtC5Me4, 1,3-Bu2tC5H3) with As4 is examined. For C5Me5 and EtC5Me4, the pentaarsaferrocene derivatives [CpRFe(η5-As5)] are formed together with [(CpRFe)3As6] and [(CpRFe)3As6{(η3-As3)Fe}], while for 1,3-Bu2tC5H3 only [(CpRFe)3As6] is formed. The reaction of [(Me5C5Fe)3As6{(η3-As3)Fe}] with Tl+ leads to [{(Me5C5Fe)3As6Fe}2(μ,η33-As3)]2+ representing an unexpected dicationic cluster.  相似文献   

4.
The thermal decompositon of a number of organo-bielemental vanadium compounds with the general formula Cp2V(ER3) (ER3 - GeEt3, SnEt3, CH2SiMe3, SeGeEt3) has been investigated in solids and in solution. The main decomposition products of Cp2V(SnEt3) are vanadocene and hexaethyldistannane. Et3GeH, Et3GeCp, Cp2V and CpV(C5H4GeEt3) are formed from Cp2V (GeET3) decomposition. Isolated CpV(C5H4GeEt3) is characterized by IR and mass spectra. The decomposition of Cp2V(CH2SiMe3) is accompanied by Me4Si, Cp2V and CpV-(C5H4CH2SiMe3) formation, the latter is identified from the mass spectrum. Triethylgermane, vanadocene, and a diselenide of vanadium are isolated on decomposition of Cp2V(SeGeEt3). Based upon the experimental data, mechanisms for the decompositon are proposed.  相似文献   

5.
Two transition-metal atoms bridged by hydrides may represent a useful structural motif for N2 activation by molecular complexes and the enzyme active site. In this study, dinuclear MoIV-FeII complexes with bridging hydrides, CpRMo(PMe3)(H)(μ-H)3FeCp* ( 2 a ; CpR=Cp*=C5Me5, 2 b ; CpR=C5Me4H), were synthesized via deprotonation of CpRMo(PMe3)H5 ( 1 a ; CpR=Cp*, 1 b ; CpR=C5Me4H) by Cp*FeN(SiMe3)2, and they were characterized by spectroscopy and crystallography. These Mo−Fe complexes reveal the shortest Mo−Fe distances ever reported (2.4005(3) Å for 2 a and 2.3952(3) Å for 2 b ), and the Mo−Fe interactions were analyzed by computational studies. Removal of the terminal Mo−H hydride in 2 a – 2 b by [Ph3C]+ in THF led to the formation of cationic THF adducts [CpRMo(PMe3)(THF)(μ-H)3FeCp*]+ ( 3 a ; CpR=Cp*, 3 b ; CpR=C5Me4H). Further reaction of 3 a with LiPPh2 gave rise to a phosphido-bridged complex Cp*Mo(PMe3)(μ-H)(μ-PPh2)FeCp* ( 4 ). A series of Mo−Fe complexes were subjected to catalytic silylation of N2 in the presence of Na and Me3SiCl, furnishing up to 129±20 equiv of N(SiMe3)3 per molecule of 2 b . Mechanism of the catalytic cycle was analyzed by DFT calculations.  相似文献   

6.
Hydrogenolysis of alkyl‐substituted cyclopentadienyl (CpR) ligated thorium tribenzyl complexes [(CpR)Th(p‐CH2‐C6H4‐Me)3] ( 1 – 6 ) afforded the first examples of molecular thorium trihydrido complexes [(CpR)Th(μ‐H)3]n (CpR=C5H2(tBu)3 or C5H2(SiMe3)3, n=5; C5Me4SiMe3, n=6; C5Me5, n=7; C5Me4H, n=8; 7 – 10 and 12 ) and [(Cp#)12Th13H40] (Cp#=C5H4SiMe3; 13 ). The nuclearity of the metal hydride clusters depends on the steric profile of the cyclopentadienyl ligands. The hydrogenolysis intermediate, tetra‐nuclear octahydrido thorium dibenzylidene complex [(Cpttt)Th(μ‐H)2]4(μ‐p‐CH‐C6H4‐Me)2 (Cpttt=C5H2(tBu)3) ( 11 ) was also isolated. All of the complexes were characterized by NMR spectroscopy and single‐crystal X‐ray analysis. Hydride positions in [(CpMe4)Th(μ‐H)3]8 (CpMe4=C5Me4H) were further precisely confirmed by single‐crystal neutron diffraction. DFT calculations strengthen the experimental assignment of the hydride positions in the complexes 7 to 12 .  相似文献   

7.
The cationic polymerization of ethyl, n-butyl and iso-butyl vinyl ether, EVE, BVE and iBVE, respectively, was efficiently conducted using bis(η5-cyclopentadienyl)dimethyl hafnium, Cp2HfMe2, or bis(η5-cyclopentadienyl)dimethyl zirconium, Cp2ZrMe2 in combination with either tris(pentafluorophenyl)borate, B(C6F5)3, or tetrakis(pentafluorophenyl)borate dimethylanilinum salt, [B(C6F5)4]?[Me2NHPh]+, as initiation systems. The evolution of polymer yield, molecular weight and molecular weight distribution with time was examined. In addition, the influence of the initiating system, the monomer and the reaction conditions on the control of the polymerization was studied. Furthermore, statistical copolymers of EVE with BVE were prepared employing Cp2HfMe2 and [B(C6F5)4]?[Me2NHPh]+ as the initiation system. The reactivity ratios were estimated using both linear graphical and non-linear methods. Structural parameters of the copolymers were obtained by calculating the dyad sequence fractions and the mean sequence length, which were derived using the monomer reactivity ratios. The glass transition temperatures, Tg, of the copolymers were measured by Differential Scanning Calorimetry, DSC, and the results were compared with predictions based on several theoretical models. The kinetics of thermal decomposition of the copolymers along with the respective homopolymers was studied by thermogravimetric analysis within the framework of the Ozawa-Flynn-Wall and Kissinger methodologies.  相似文献   

8.
A series of nickel complexes, including Ni(acac)2, (C5H5)Ni(η3‐allyl), and [NiMe4Li2(THF)2]2, that were activated with modified methylaluminoxane (MMAO) exhibited high catalytic activity for the polymerization of methyl methacrylate (MMA) but showed no catalytic activity for the polymerization of ethylene and 1‐olefins. The resulting polymers exhibited rather broad molecular weight distributions and low syndiotacticities. In contrast to these initiators, the metallocene complexes (C5H5)2Ni, (C5Me5)2Ni, (Ind)2Ni, and (Me3SiC5H4)2Ni provided narrower molecular weight distributions at 60 °C when these initiator were activated with MMAO. Half‐metallocene complexes such as (C5H5)NiCl(PPh3), (C5Me5)NiCl(PPh3), and (Ind)NiCl(PPh3) produced poly(methyl methacrylate) (PMMA) with much narrower molecular weight distributions when the polymerization was carried out at 0 °C. Ni[1,3‐(CF3)2‐acac]2 generated PMMA with high syndiotacticity. The NiR(acac)(PPh3) complexes (R = Me or Et) revealed high selectivity in the polymerization of isoprene that produced 1,2‐/3,4‐polymer at 0 °C exclusively, whereas the polymerization at 60 °C resulted in the formation of cis‐1,4‐rich polymers. The polymerization of ethylene with Ni(1,3‐tBu2‐acac)2 and Ni[1,3‐(CF3)2‐acac]2 generated oligo‐ethylene with moderate catalytic activity, whereas the reaction of ethylene with Ni(acac)2/MMAO produced high molecular weight polyethylene. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4764–4775, 2000  相似文献   

9.
The relative rate constants of the addition of the C6H5CH2 radical to unsaturated compounds CH2=CHX (X = C4H9, SiMe3, CF3, CO2Me, CN) were determined under the conditions of initiation by the Fe(CO)5 + DMF system or by benzoyl peroxide. Depending on the values of the relative addition rate constants, the monomers can be arranged into the following series (X): CF3C4H932Me5 + DMF system, the addition stage proceeds by a free radical mechanism.A. N. Nesmeyanov Institute of Heteroorganic Compounds, Russian Academy of Sciences, 117813 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 9, pp. 2017–2022, September, 1992.  相似文献   

10.
An idea of making a ferrocene/fullerene hybrid, “bucky ferrocene”, has intrigued chemists for some time, but the compounds remained to be hypothetical. The synthesis of such hybrid molecules as Fe(C60Me5)Cp, Ru(C60Me5)Cp and Fe(C70Me3)Cp as well as their functionalized derivatives from [60] and [70]fullerenes has been achieved in recent years. With their esthetically pleasing structures and the dual character of metallocene and graphite, these molecules may stimulate the interest of both chemists and non-chemists.  相似文献   

11.
The oligomerization and polymerization of 1‐pentene using Cp2ZrCl2, Cp2HfCl2, [(CH3)5C5]2ZrCl2, rac‐[C2H4(Ind)2]ZrCl2, [(CH3)2Si(Ind)2]ZrCl2, (CH3)2Si(2‐methylbenz[e]indenyl)2ZrCl2, Cp2ZrCl{O(Me)CW(CO)5}, Cp2ZrCl(OMe) and methylaluminoxane (MAO) has been studied. The degree of polymerization was highly dependent on the metallocene catalyst. Oligomers ranging from the dimer of 1‐pentene to polymers of poly‐1‐pentene with a molar mass Mw = 149000 g/mol were formed. Cp2ZrCl{O(Me)CW(CO)5} is a new highly active catalyst for the oligomerization of 1‐pentene to low molecular weight products. The activity decreases in the order Cp2ZrCl{O(Me)CW(CO)5} > Cp2ZrCl2 > Cp2ZrCl(OMe). Furthermore, poly‐1‐olefins ranging from poly‐1‐pentene to poly‐1‐octadecene were synthesized with (CH3)2Si(2‐methyl‐benz[e]indenyl)2ZrCl2 and methylaluminoxane (MAO) at different temperatures. The temperature dependence of the molar mass can be described by a common exponential decay function irrespective of the investigated monomer.  相似文献   

12.
Ethylene/1-hexene copolymers produced with MAO-activated binary metallocene catalysts, such as combinations Cp2ZrCl2 + (Me5Cp)2ZrCl2, (Ind-H4)2ZrCl2 + (Me5Cp)2ZrCl2, Cp2ZrCl2 + Cp2TiCl2, etc., contain three types of components. Two of the components can be attributed to active centers derived from each individual metallocene complex, and one or two materials are produced with different types of active center. Some of the binary catalysts generate the three components in comparable proportions, whereas other catalysts produce copolymers with one dominant component, which does not resemble the copolymers produced with the individual complexes. A mechanism is proposed for the formation of the “new” copolymer materials.  相似文献   

13.
The 15 valence-electron iron(I) complex [CpArFe(IiPr2Me2)] ( 1 , CpAr=C5(C6H4-4-Et)5; IiPr2Me2=1,3-diisopropyl-4,5-dimethylimidazolin-2-ylidene) was synthesized in high yield from the FeII precursor [CpArFe(μ-Br)]2. 57Fe Mössbauer and EPR spectroscopic data, magnetic measurements, and ab initio ligand-field calculations indicate an S= 3/2 ground state with a large negative zero-field splitting. As a consequence, 1 features magnetic anisotropy with an effective spin-reversal barrier of Ueff=64 cm−1. Moreover, 1 catalyzes the dehydrogenation of N,N-dimethylamine–borane, affording tetramethyl-1,3-diaza-2,4-diboretane under mild conditions.  相似文献   

14.
Density functional theory was used to study gas-phase reactions between the Cp2*ZrMe+ cations, where Cp* = C5H5 (1), Me5Cp = C5Me5 (2), and Flu = C13H9 (3), and the ethylene molecule, Cp2*ZrMe+ + C2H4 → Cp2*ZrPr+ → Cp2*ZrAllyl+ + H2. The reactivity of the Cp2*ZrMe+ cations with respect to the ethylene molecule decreased in the series 1 > 32. Substitution in the Cp ring decreased the reactivity of the Cp2*ZrMe+ cations toward ethylene, in agreement with the experimental data on the comparative reactivities of complexes 1 and 3. The two main energy barriers along the reaction path (the formation of the C-C bond leading to the primary product Cp2*ZrPr+ and hydride shift leading to the secondary product Cp2*Zr(H2)Allyl+) vary in opposite directions in the series of the compounds studied. For Flu (3), these barriers are close to each other, and for the other compounds, the formation of the C-C bond requires the overcoming of a higher energy barrier. A comparison of the results obtained with the data on the activity of zirconocene catalysts in real catalytic systems for the polymerization of ethylene led us to conclude that the properties of the catalytic center changed drastically in the passage from the model reaction in the gas phase to real catalytic systems.  相似文献   

15.
Reaction of Cp2Ta(H)L (L = C3H6, C4H8, C5H10 and C5H8) with CO under mild conditions gives alkyltantalum carbonyl complexes Cp2Ta(CO)R (R = C3H7, C4H9, C5H11 and C5H9, respectively). Depending upon the position of the olefin relative to the hydride ligand in the hydride-olefin complex, Cp2Ta(CO)H is also formed during the carbonylation reaction. Reduction of Cp2TaCl2 by potassium or t-BuMgCl under one atmosphere of CO affords the very stable compound Cp2Ta(CO)Cl in moderate yields. Reaction of Cp2Ta(CO)Cl with RLi or RMgX does not give the Cp2Ta(CO)R complex.  相似文献   

16.
This paper presents an extensive study of the polymerization of MMA with borohydrido lanthanide complexes for the first time. Catalytic systems are made from a lanthanide derivative bearing zero one, or two bulky ligands: substituted cyclopentadienyl (Cp*′ = C5Me4nPr, Cp4i = C5HiPr4, CpPh3 = H2C5Ph3‐1,2,4), and/or diketiminate ([(p‐tol)NN] = [(p‐CH3C6H4)N(CH3)C]2CH), in the presence of variable quantities of alkylating agent. With BuLi in apolar medium, highly isotactic polymer (up to 95.6%) is formed. In THF, syndiotactic‐rich PMMA is obtained whatever the nature of the co‐catalyst (BuLi or MgnBu2). The presence of an electron‐withdrawing ligand such as CpPh3 allows high syndioregularity, up to 81.8% at 0 °C, together with the highest conversion. There is quite good concordance between calculated and experimental molecular data in THF. Divalent Cp*′2SmII(THF) and (CpPh3)2SmII(THF) are active as single‐component initiators; the former affords PMMA 88% syndiotactic at 0 °C. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

17.
The photoacoustic spectrum (PAS) was measured in the near IR region (1000 to 2600 nm) for organic compounds (C6H6, C6D6, C7H8, and C6H12) and organometallic compounds (Cp2Fe, Cp2Fe2(CO)4, Cp4Fe4(CO)4, Cp4Fe4S4, C6H6Cr(CO)3, CpCo(C4Ph4) and CpCo(CO)2). Band assignments were made by comparison to the infrared spectra. The bands were assigned as the CH overtone stretch and combinations of CH and other IR fundamentals. These bands provide fingerprint spectra for these compounds.  相似文献   

18.
The “borohydride/alkyl” (B/A) route initially reported for isoprene has been applied successfully to the polymerization of styrene. This method provides via an in situ approach an interesting tool for the assessment of the influence of a ligand on the performance of half-lanthanidocene catalysts. All systems lead to well-controlled oligomerization/polymerization processes. This method is thus a convenient tool for the controlled polymerization of styrene starting from a common trisborohydride precursor and commercial ligands. The influence of the nature of several ligands on the activity could be established, with trends corresponding to those obtained starting from the isolated precursors: HCpHCpPh3>HCp*(Cp=C5H5,CpPh3=1,2,4-Ph3C5H2,Cp*=C5Me5). These results suggest an influence of the electron donating ability of the ligand rather than steric requirements.  相似文献   

19.
The reaction of different metallocene fragments [Cp2M] (Cp=η5‐cyclopentadienyl, M=Ti, Zr) with diferrocenylacetylene and 1,4‐diferrocenylbuta‐1,3‐diyne is described. The titanocene complexes form the highly strained three‐ and five‐membered ring systems [Cp2Ti(η2‐FcC2Fc)] ( 1 ) and [Cp2Ti(η4‐FcC4Fc)] ( 2 ) (Fc=[Fe(η5‐C5H4)(η5‐C5H5)]) by addition of the appropriate alkyne or diyne to Cp2Ti. Zirconocene precursors react with diferrocenyl‐ and ferrocenylphenylacetylene under C? C bond coupling to yield the metallacyclopentadienes [Cp2Zr(C4Fc4)] ( 3 ) and [Cp2Zr(C4Fc2Ph2)] ( 5 ), respectively. The exchange of the zirconocene unit in 3 by hydrogen atoms opens the route to the super‐crowded ferrocenyl‐substituted compound tetraferrocenylbutadiene ( 4 ). On the other hand, the reaction of 1,4‐diferrocenylbuta‐1,3‐diyne with zirconocene complexes afforded a cleavage of the central C? C bond, and thus, dinuclear [{Cp2Zr(μ‐η12‐C?CFc)}2] ( 6 ) that consists of two zirconocene acetylide groups was formed. Most of the complexes were characterized by single‐crystal X‐ray crystallography, showing attractive multinuclear molecules. The redox properties of 3 , 5 , and 6 were studied by cyclic voltammetry. Upon oxidation to 3 n+, 5 n+, and 6 n+ (n=1–3), decomposition occured with in situ formation of new species. The follow‐up products from 3 and 5 possess two or four reversible redox events pointing to butadiene‐based molecules. However, the dinuclear complex 6 afforded ethynylferrocene under the measurement conditions.  相似文献   

20.
Reactions of Cp2TaCl2 with RMgCl (R = n-C3H7, i-C3H7, n-C4H9, s-C4H9, n-C5H11 and C5H9) give tantalum hydride π-olefin complexes Cp2Ta(H)L (L = C3H6, C4H8, C5H10 and C5H8). Two isomers of Cp2Ta(H)C3H6 were obtained. The complexes are useful starting materials for the synthesis of other tantalum hydride species, e.g. Cp2Ta(H)PEt3 and Cp2TaH3.  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