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 905 毫秒
1.
Ring‐opening metathesis polymerization of dicyclopentadiene catalyzed by TiCl4 · 2L/CH3Li system [where L is tetrahydropyran (1), dioxane, 2,5‐dimethylfuran, or tetrahydrofurfyl alcohol] is reported. The obtained polymer was characterized by IR and 1H‐NMR. These catalytic systems effectively promoted the polymerization reaction. Seven influencing factors are discussed. When the aging temperature was 0°C, the aging time was 90 min, the polymerization temperature was 60°C, Li/Ti was 1.5–2, and the monomer/catalyst molar ratio ranged between 30 and 50, the polymerization reaction catalyzed by complex 1 yielded better results within a shorter period of time. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 81: 662–666, 2001  相似文献   

2.
The polymerization of styrene with a series of bispyrazolylimine dinickel (II) complexes of bis‐2‐(C3HN2(R1)2‐3,5)(C(R2) = N(C6H3(CH3)2‐2,6)Ni2Br4 (complex 1 : R1 = CH3, R2 = Ph; complex 2 : R1 = CH3, R2 = 2,4,6‐trimethylphenyl; complex 3 : R1 = R2 = Ph; complex 4 : R1 = Ph, R2 = 2,4,6‐trimethylphenyl) in the presence of methylaluminoxane (MAO) was studied. The influences of polymerization parameters such as polymerization temperature, Al/Ni molar ratio, reaction time, and catalyst concentration on catalytic activity and molecular weight of the polystyrene were investigated in detail. The influence of the bulkiness of the substituents on polymerization activity was also studied. All of the four catalytic systems exhibited high activity (up to 10.50 × 105 gPS/(mol Ni h)) for styrene polymerization and provide polystyrene with moderate to low molecular weights (Mw = 4.76 × 104–0.71 × 104 g/mol) and narrower molecular weight distributions about 2. The obtained polystyrene was characterized by means of FTIR, 1H‐NMR, and 13C‐NMR techniques. The results indicated that the polystyrene was atactic polymer. The analysis of the end groups of polystyrene indicated that styrene polymerization with bispyrazolylimine dinickel complexes/MAO catalytic systems proceeded through a coordination mechanism. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011.  相似文献   

3.
A series of nonbridged (cyclopentadienyl) (aryloxy)titanium(IV) complexes of the type, (η5‐Cp′)(OAr)TiCl2 [OAr = O‐2,4,6‐tBu3C6H2 and Cp′ = Me5C5 ( 1 ), Me4PhC5 ( 2 ), and 1,2‐Ph2‐4‐MeC5H2 ( 3 )], were prepared and used for the copolymerization of ethylene with α‐olefins (e.g., 1‐hexene, 1‐octene, and 1‐octadecene) in presence of AliBu3 and Ph3CB(C6F5)4 (TIBA/B). The effect of the catalyst structure, comonomer, and reaction conditions on the catalytic activity, comonomer incorporation, and molecular weight of the produced copolymers was examined. The substituents on the cyclopentadienyl group of the ligand in 1 – 3 play an important role in the catalytic activity and comonomer incorporation. The 1 /TIBA/B catalyst system exhibits the highest catalytic activity, while the 3 /TIBA/B catalyst system yields copolymers with the highest comonomer incorporation under the same conditions. The reactivity ratio product values are smaller than those by ordinary metallocene type, which indicates that the copolymerization of ethylene with 1‐hexene, 1‐octene, and 1‐octadecene by the 1–3/ TIBA/B catalyst systems does not proceed in a random manner. © 2011 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

4.
Microporous polyurethane elastomer, based on 4,4′‐diphenylmethane diisocyanate and polyester polyol Bayflex 2003 (Bayer AG), was degraded by phosphoric acid esters  (CH3CH2O)3P(O) and (ClCH2CH2O)3P(O) at 180°C. Structure of degraded products was investigated by means of 1H‐, 13C‐, and 31P‐NMR spectroscopy. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 76: 886–893, 2000  相似文献   

5.
The synthesis of a new family of single‐ion conducting random copolymers bearing polyhedral boron anions is reported. For this purpose two novel ionic monomers, namely [B12H11(OCH2CH2)2OC(?O)C(CH3)?CH2]2?[(C4H9)4N+]2 and [8‐(OCH2CH2)2OC(?O)C(CH3)?CH2‐3,3′‐Co(1,2‐C2B9H10)(1′,2′‐C2B9H11)]?K+, having methacrylate function, diethylene glycol bridge and closo‐dodecaborate or cobalt bis(1,2‐dicarbollide) anions were designed. Such monomers differ from previously reported ones by (i) chemically attached highly delocalized boron anions, by (ii) valency of the anion (divalent anion and monovalent one) and by (iii) the presence of oxyethylene flexible spacer between the methacrylate group and bonded anion. Their free radical copolymerization with poly(ethylene glycol) methyl ether methacrylate and subsequent ion exchange provided lithium‐ion conducting polyelectrolytes showing low glass transition temperature (?53 to ?49 °C), ionic conductivity up to 9.1 × 10?7 S cm?1, lithium transference number up to 0.61 (70 °C) and electrochemical stability up to 4.1 V versus Li+/Li (70 °C). The incorporation of propylene carbonate (20–40 wt%) into the copolymers resulted in the enhancement of their ionic conductivity by one order of magnitude and significantly increased their electrochemical stability up to 4.7 V versus Li+/Li (70 °C). © 2019 Society of Chemical Industry  相似文献   

6.
Summary: A comparison of PP qualities, which are produced with two different polymerization techniques–gas phase(GP) and liquid pool (LP)–under precise control of the polymerization temperature (70 °C) and pressure (GP = 25 bar, LP < 60 bar) using identical Ziegler‐Natta catalyst (TiCl4/phthalate/MgCl2 + TEA/silane), is presented. A series of homopolymer PP in a wide MW range from 100 000 to 1 600 000 g · mol?1 was polymerized. During polymerization all samples were characterized exactly by their Rp‐profil. The effect of hydrogen on the initial reaction rate and on MW and MWD was analyzed on the basis of this so‐called kinetic fingerprint. The results showed that the polymerization rate reached a maximum for LP, of about 150 kg · gcat?1 · h?1, in contrast to GP with a maximum of Rp,0 = 45 kg · gcat?1 · h?1. Analysis was carried out by means of GPC, SEM, DSC, platte‐platte rheometer, and WAXS. The results first showed that the MWD of LP PP is narrower (PD ~ 6.8) than for the GP PP (PD ~ 8), polymerized in two steps. An SEM study of the powder particle shows the typical dent surface morphology of polymers using Ziegler‐Natta catalysts for polymerization. WAXS and DSC analysis demonstrated that almost only the α‐modification of crystalline structure exists and that the crystallinity becomes considerably higher after solidification from melt. Furthermore, it was found that the crystallite size distribution depends on the polymerization technique. Rheological studies indicate that GP PP behaves more elastically. To summarize, it is shown that PP produced with the LP polymerization technique is more homogenous and of high quality.

Particle geometry of gas phase and liquid pool polymerized PP powder observed by SEM (PP‐L0).  相似文献   


7.
Summary Halogen-free polyisobutylene (PIB) was synthesized by in situ methylation of living PIB using dimethylzinc. Quantitative methylation of living PIB was achieved within 60 min using a ratio of [(CH3)2Zn]/[TiCl4]0= 1 without any side reactions. Under similar conditions, living PIB capped with 1,1-diphenylethylene (PIB-DPE+) yielded close to 1:1 mixture of methoxy- and methyl-functionality. By using the ratio of [(CH3)2Zn]/[TiCl4]0≥ 3, however, quantitative methylation of PIB-DPE+ could be achieved in 120 min without any side reactions as confirmed by spectroscopic and chromatographic analyses. Received: 1 February 2000/Revised version: 23 April 2000/Accepted: 23 April 2000  相似文献   

8.
A series of novel cationic gemini surfactants [CnH2n+1–O–CH2–CH(OH)–CH2–N+(CH3)2–(CH2)2]2·2Br? [ 3a (n = 12), 3b (n = 14) and 3c (n = 16)] having a 2‐hydroxy‐1,3‐oxypropylene group [?CH2–CH(OH)–CH2–O–] in the hydrophobic chain have been synthesized and characterized. Their water solubility, surface activity, foaming properties, and antibacterial activity have been examined. The critical micelle concentration (CMC) values of the novel cationic gemini surfactants are one to two orders of magnitude smaller than those of the corresponding monomeric surfactants. Furthermore, the novel cationic gemini surfactants have better water solubility and surface activity than the comparable [CnH2n+1–N+(CH3)2–(CH2)2]2·2Br? (n‐4‐n) geminis. The novel cationic gemini surfactants 3a and 3b also exhibit good foaming properties and show good antibacterial and antifungal activities.  相似文献   

9.
The reactions of 2-(((2-hydroxy-3-methoxyphenyl)methylene)amino)-2-(hydroxymethyl)-1,3-propanediol (H4L) and Mn(ClO4)2·6H2O or Co(SCN)2·3H2O in the presence of triethylamine in methanol led to the formation of two new complexes [MnΙΙΙ4(HL)2(H2L)2(CH3OH)4]·4CH3OH·(ClO4)2 (1) and [CoΙΙCoΙΙΙ(H2L)2(CH3OH)(SCN)]·1.5CH3OH·1.5H2O (2), respectively. According to structural data and magnetic properties tetranuclear complex 1 contains four homo-valence manganese (ΙΙΙ) atoms, while in the binuclear complex 2 composed of hetero-valence bi- and trivalent cobalt (ΙΙ, ΙΙΙ) atoms. Weak antiferromagnetic exchange interactions between neighboring manganese ions in 1 have place. χMT for 2 was fitted using a model of isolated cobalt (ΙΙ) ion with zero-field splitting parameters and the study confirms its mixed valence CoΙΙ/CoΙΙΙ nature. No slow magnetic relaxation effects were observed for both complexes in the absence of an applied dc magnetic field.  相似文献   

10.
Treatment of (tBu3SiNH)(tBu3SiN=)2WH ( 1 -H) with small alkyl anions (RM) afforded tungsten alkyl hydride anions [(tBu3SiNH)(tBu3SiN=)2HWR)]M ( 3 -(R)M: R=CH3, M=Li; R=nBu, M=Li; R=neoPe, M=Li; R=CH2Ph (Bn), M=K (two isomers); R=CCH, M=Na; R=CH=CH2 (Vy), M=Li). The saturated alkyl anions 3 -(R)M ( 3 -(R)M: R=CH3, M=Li; R=nBu, M=Li; R=neoPe, M=Li; R=CH2Ph (Bn), M=K) degraded via apparent 1,2-RH-elimination to produce the known [(tBu3SiN=)3WH]M ( 2 -HM), but the acetylide ( 3 -(C2H)Na) and vinyl ( 3 -(Vy)Li) anions converted to their hydrogenated isomers, [(tBu3SiN=)3WVy]Na ( 2 -VyNa) and [(tBu3SiN=)3WEt]Li ( 2 -EtLi), respectively. The structure of 3 -(nBu)Li is reported, and a discussion of tungsten-hydride coupling constants in tBu3SiX-ligated (X=O, NH, N) systems is given.  相似文献   

11.
The formation of CH-type catalysts has been investigated by high-resolution and solid-state NMR. These catalysts are prepared from a soluble MgCl2 and 2-ethyl-1-hexanol adduct (MgCl2·3EH) by reaction with phthalic anhydride (PA) to form dioctylphthalate (DOP) and then with TiCl4 in the presence of di-i-butylphthalate (BP). In the model systems MgCl2·3EH/PA, MgCl2/BP, and MgCl2/TiCl4/BP, the ester is complexed with MgCl2 and /or TiCl4 in two or more states. Only single-ester C?O and OCH2 resonances are seen in TiCl4/BP, probably due to exchange of ester coordinations. CH-catalysts prepared by three different procedures exhibit a single mode of bonding for the ester. The chemical shift values are consistent for ester complexed with MgCl2. The most active and stereoselective catalyst has the most shielded chemical shift values for the C?O and —OCH2— carbons, shortest TH1 and TH1p, and longest TCH relaxation times. These parameters change monotonically with the decrease of activity and stereoselectivity of the catalyst preparation. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Polyethylene glycol 400 was reacted with CH2Cl2 in the presence of KOH to form oxymethylene-linked chains. The method gave a high yield of colourless high-molecular-weight elastomer. The ionic conductivity of a mixture of the polymer with LiCF3SO3 ([O]/[Li] = 25) was about 5 × 10?5S cm?1 at 25°C.  相似文献   

13.
Optimization of the fermentation media for maximization of surfactin production was carried out. The carbon source (glucose), the nitrogen source (ammonium nitrate) and the mineral salts ferrous and manganous sulphates were the critical components of the medium optimized. A 24 full factorial central composite experimental design followed by multi-stage Monte-Carlo optimization was used in the design of experiments and in the analysis of results. This procedure limited the number of actual experiments performed while allowing for possible interactions between the four components. The optimum values for the tested variables for the maximal production of surfactin were (in g dm−3): glucose = 36·5; NH4NO3 = 4·5; FeSO4 = 4×10−3 and MnSO4 = 27·5 ×10−2. Relative surfactant concentrations were expressed as the reciprocal of the critical micelle concentration (CMC−1) and the maximum predicted yield of surfactin in terms of CMC−1 was 45·5. © 1997 SCI.  相似文献   

14.
Reactions of bis(benzoylacetone)-1,3-diiminopropan-2-ol (abbreviated as H3L) with zinc salts in the presence of triethylamine afforded the compounds [Zn4(HL)4]·4CH3CN (1·4CH3CN) and [Zn8L4(OH)4]·2CH3CN (2·2CH3CN). Further reaction of 1 with Ni(CH3COO)2·4H2O gave the heteronuclear species [Zn2Ni2(L)2(CH3O)2(CH3OH)2] (3). The crystal structures of 1·4CH3CN, 2·2CH3CN and 3 were determined by the X-ray diffraction method.  相似文献   

15.
The separation of CO2/CH4 is reported in detail by using zeolitic imidazolate framework (ZIF-8) membrane which was prepared on 3-aminopropyltriethoxysilane modified Al2O3 tube through microwave heating synthesis. Attributed to the preferential adsorption affinity of CO2 over CH4 and a narrow pore window of 0.34 nm, the ZIF-8 membrane shows high separation performances for the separation of CO2/CH4 mixtures. For the separation of equimolar CO2/CH4 mixture at 100°C and 2 bar feed (1 bar permeate) pressure, a CO2 permeance of 1.02 × 10?8 mol/m2· s· Pa and a CO2/CH4 selectivity of 6.8 are obtained, which is promising for CO2 separation.  相似文献   

16.
CO2 reforming, oxidative conversion and simultaneous oxidative conversion and CO2 or steam reforming of methane to syngas (CO and H2) over NiO–CoO–MgO (Co: Ni: Mg=0·5: 0·5:1·0) solid solution at 700–850°C and high space velocity (5·1×105 cm3 g−1 h−1 for oxidative conversion and 4·5×104 cm3 g−1 h−1 for oxy-steam or oxy-CO2 reforming) for different CH4/O2 (1·8–8·0) and CH4/CO2 or H2O (1·5–8·4) ratios have been thoroughly investigated. Because of the replacement of 50 mol% of the NiO by CoO in NiO–MgO (Ni/Mg=1·0), the performance of the catalyst in the methane to syngas conversion process is improved; the carbon formation on the catalyst is drastically reduced. The CoO–NiO–MgO catalyst shows high methane conversion activity (methane conversion >80%) and high selectivity for both CO and H2 in the oxy-CO2 reforming and oxy-steam reforming processes at ⩾800°C. The oxy-steam or CO2 reforming process involves the coupling of the exothermic oxidative conversion and endothermic CO2 or steam reforming reactions, making these processes highly energy efficient and also safe to operate. These processes can be made thermoneutral or mildly exothermic or mildly endothermic by manipulating the process conditions (viz. temperature and/or CH4/O2 ratio in the feed). © 1998 Society of Chemistry Industry  相似文献   

17.
We report a novel, high yield one-step synthesis of water stable and soluble titanocene dichloride dihydrochloride salts from the direct reaction of the neutral amino-substituted cyclopentadienes with TiCl4. The following novel complexes have been synthesised: C5H4(CH2)2N(CH2)5]2TiCl2 (5), [C5H4CH(CH2)4NMe]2TiCl2 (6), [C5H4(CH2)2N(CH2)5]2TiCl2·2HCl (7), [C5H4CH(CH2)4]2NMeTiCl2·2HCl (8), [C5H4(CH2)2N(CH2)5]2TiMe2 (9).  相似文献   

18.
In this study, the effects of 1-Ethyl-3-methylimidazolium tetrafluoroborate ionic liquid on CO2/CH4 separation performance of symmetric polysulfone membranes are investigated. Pure polysulfone membrane and ionic liquid-containing membranes are characterized. Field emission scanning electron microscopy (FE-SEM) is used to analyze surface morphology and thickness of the fabricated membranes. Energy dispersive spectroscopy (EDS) and elemental mapping, Fourier transform infrared (FTIR), thermal gravimetric (TGA), X-ray diffraction (XRD) and Tensile strength analyses are also conducted to characterize the prepared membranes. CO2/CH4 separation performance of the membranes are measured twice at 0.3 MPa and room temperature (25 °C). Permeability measurements confirm that increasing ionic liquid content in polymer-ionic liquid membranes leads to a growth in CO2 permeation and CO2/CH4 selectivity due to high affinity of the ionic liquid to carbon dioxide. CO2 permeation significantly increases from 4.3 Barrer (1 Barrer=10-10 cm3(STP)·cm·cm-2·s-1·cmHg-1, 1cmHg=1.333kPa) for the pure polymer membrane to 601.9 Barrer for the 30 wt% ionic liquid membrane. Also, selectivity of this membrane is improved from 8.2 to 25.8. mixed gas tests are implemented to investigate gases interaction. The results showed, the disruptive effect of CH4 molecules for CO2 permeation lead to selectivity decrement compare to pure gas test. The fabricated membranes with high ionic liquid content in this study are promising materials for industrial CO2/CH4 separation membranes.  相似文献   

19.
Linear asymmetrical poly(propylene oxide) was synthesized through four‐step reactions: selective benzylation, alcohol exchange reaction, propylene oxide anionic polymerization, debenzylation. One terminal of the asymmetrical polymer chains is alcohol hydroxyl and the other is phenol hydroxyl. It was characterized with infrared (IR) and 1H Nuclear Magnetic Resonance (1H‐NMR). Peaks at 1.11, 3.38, and 3.53 ppm were attributed to side groups (? OCH2CH(CH3)? ), backbone units (? OCH2CH(CH3)? ) and (? OCH2CH(CH3)? ) of poly(propylene oxide), respectively. Molecular weight and molecular weight distribution were measured with 1H‐NMR and laser light scattering (LLS), which showed that the linear asymmetrical poly(propylene oxide) was mono‐disperse (PDI = 1.02–1.07). Then, its carbamate reaction with phenyl isocyanate was studied; the reaction rate constants for phenol hydroxyl and alcohol hydroxyl of poly(propylene oxide) were k1 = 0.209 mol L?1 min?1 and k2 = 0.051 mol L?1 min?1. There was a great reactivity difference for two types of hydroxyls in asymmetrical poly(propylene oxide), contrasting to the single carbamate reaction rate constant of symmetrical poly(propylene oxide) (k3 = 0.049 mol L?1 min?1). © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

20.
Alkali metal (viz. Li, Na, K, Rb and Cs) promoted MgO catalysts (with an alkali metal/Mg ratio of 0·1) calcined at 750°C have been compared for their surface properties (viz. surface area, morphology, acidity and acid strength distribution, basicity and base strength distribution, etc.) and catalytic activity/selectivity in the oxidative coupling of methane (OCM) to C2-hydrocarbons at different temperatures (700–750°C), CH4/O2 ratios (4·0 and 8·0) in feed, and space velocities (10320 cm3 g−1 h−1). The surface and catalytic properties of alkali metal promoted MgO catalysts are found to be strongly influenced by the alkali metal promoter and the calcination temperature of the catalysts. A close relationship between the surface density of strong basic sites and the rate of C2-hydrocarbons formation per unit surface area of the catalysts has been observed. Among the catalysts calcined at 750°C, the best performance in the OCM is shown by Li–MgO (at 750°C). © 1997 SCI.  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