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.
In the soapfree emulsion polymerization of a methyl methacrylate-K2S2O8-CaSO3-H2O system, the polymerization rate, average molecular weight of polymer, particle size and particle concentration would vary with the concentration of CaSO3. It was shown that when the concentration of CaSO3 was well below the saturation concentration (3 × 10?4mol/litre H2O), the polymerization rate was higher than that of the system not containing CaSO3. On the other hand, when the concentration of CaSO3 was above the saturation concentration, the polymerization rate at the latter stage was lower than that of the system not containing CaSO3 within our experimental conditions. The molecular weight of polymer was measured by Gel Permeation Chromatography (GPC). It decreased initially and then increased due to the gel effect over the entire course of polymerization. The size of the polymer particles was measured by both photo correlation spectroscopy (PCS) and transmission electron microscopy (TEM), The reaction mechanism was studied according to the above observation. The mechanical property of poly(methyl methacrylate)-CaSO3 composite obtained from soapfree emulsion polymerization was tested and compared with that obtained from mechanical blending.  相似文献   

2.
The butadiene polymerization in toluene at 25°C on VOCl3? (n‐C4H9)Mg(iso‐C8H17) catalytic system was investigated. The kinetic parameters of polymerization and molecular characteristics of polybutadiene were determined. It was shown that substitution of traditional organoaluminum cocatalysts in trans‐regulating vanadium systems does not have an effect on its stereospecificity, but significantly influences on the active centers reaction ability. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 596–600, 2003  相似文献   

3.
Copolymerization of ethylene with 1‐octadecene was studied using [η51‐C5Me4‐4‐R1‐6‐R‐C6H2O]TiCl2 [R1 = tBu (1), H (2, 3, 4); R = tBu (1, 2), Me (3), Ph (4)] as catalysts in the presence of Al(i‐Bu)3 and [Ph3C][B(C6F5)4]. The effect of the concentration of comonomer in the feed and Al/Ti molar ratio on the catalytic activity and molecular weight of the resultant copolymer were investigated. The substituents on the phenyl ring of the ligand affect considerably both the catalytic activity and comonomer incorporation. The 1 /Al(i‐Bu)3/[Ph3C][B(C6F5)4] catalyst system exhibits the highest catalytic activity and produces copolymers with the highest molecular weight, while the 2 /Al(i‐Bu)3/[Ph3C][B(C6F5)4] catalyst system gives copolymers with the highest comonomer incorporation under similar conditions. © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2011  相似文献   

4.
The polymerization of butadiene in toluene using Co(acac)3–(i-Bu)3Al–H2O catalyst system was studied. Presented are the effects of the addition order, aging time, and composition of catalysts on rates, polymer microstructure, and molecular weights. The polymerization was found to be initiated by the Co(acac)3-hydrolized aluminum alkyl complex. The chain propagation proceeds according to a first-order reaction with respect to monomer and active species and is a strong function of Al/H2O with an optimum ratio of 1.0, but independent of Al/Co. The nature of polymerization seems to change as Al/H2O increases from less than 1 to greater than 1. Transfer reaction is significant. From the kinetic data it was found that the termination reaction is most likely to be by combination.  相似文献   

5.
AlEt3Cl was modified with TLTTP (trilauryltrithiophosphite) in the catalyst system consisting of TiCl3 and AlEt2Cl. The effects of TLTTP on the polymerization of propylene were studied in comparison with those of alkyl homologues of TLTTP. The catalytic behavior of the TiCl3–AlEt2Cl-TLTTP catalyst system in the polymerization of propylene was also studied in comparison with that of the TiCl3–AlEt2Cl catalyst system. In the study of the effect of various alkylthiophosphites added, it is found that the bulkiness of the alkyl group affects the rate of propylene polymerization and the stereoregularity of the resultant polymers. The TiCl3–AlEt2Cl–TLTTP catalyst system gave different catalytic behavior in the propylene polymerization from that of the unmodified conventional catalyst system (TiCl3–AlEt2Cl). These effects of TLTTP were considered to be due to the bulkiness of the alkyl groups attached to the phosphorous atom and the higher reactivity to TiCl3 of the modified AlEt2Cl than of the unmodified AlEt2Cl.  相似文献   

6.
A study of nitrous oxide (N2O) reduction with methane (CH4) and propene (C3H6) in the presence of oxygen (5%) over Ag/Al2O3, Rh/Al2O3 and Ag–Rh/Al2O3 catalysts, with Ag and Rh loadings of 5 wt% and 0.05 wt% respectively, has been performed. From the results, it was observed that the Ag–Rh bimetallic catalyst was the most active for both nitrous oxide removal (more than 95%) and hydrocarbon oxidation. This high activity seems to be connected with a synergistic effect between Ag and Rh. The findings from X‐ray diffraction and X‐ray photoelectron spectroscopy studies showed also, that there were no strong interactions (eg alloying) between Ag and Rh. Copyright © 2005 Society of Chemical Industry  相似文献   

7.
La2O3–Nb2O5–Al2O3 high‐refractive‐index glasses were fabricated by containerless processing, and the glass‐forming region was determined. The thermal stability, density, optical transmittance, and the refractive index dispersion of these glasses were investigated. All the glasses were colorless and transparent in the visible to near infrared (NIR) region and had high refractive index with low wavelength dispersion. Some of these glasses were found to have significantly high glass‐forming ability. These results indicate that the ternary glasses are suitable for optical applications in the visible to NIR region. The effects of the substitution of Al2O3 for Nb2O5 on optical properties were discussed on the basis of the Drude–Voigt equation. It was suggested that the substitution of Al2O3 for Nb2O5 increased the molecular density and suppressed a decrease in the refractive index, even when both the average oscillator strength and inherent absorption wavelength decreased in La2O3–Nb2O5–Al2O3 glasses. These results are helpful for designing new optical glasses controlled to have a higher refractive index and lower wavelength dispersion.  相似文献   

8.
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   

9.
S,S‐Acetals are smoothly deprotected with air in the presence of a catalytic amount of Bi(NO3)3⋅5 H2O (1–50 mol %) under ambient conditions to regenerate the original carbonyl compounds in good to excellent yield. This mild, simple, and environmentally benign system is successfully applied to the deprotection of S,O‐ and O,O‐acetals and is compatible with various functional groups. From the mechanistic study of the reaction, the catalytic cycle is considered to be composed of the following four steps: (1) the nitrososulfonium ion of the S,S‐acetal is formed by attack of nitrosonium ion (NO +) generated from Bi(NO3)3⋅5 H2O through the equilibrium with NO2, (2) the nitrososulfonium ion is hydrolyzed to afford the hemithioacetal and thionitrite, (3) the hemithioacetal collapses to the original carbonyl compound and thiol, which is oxidized by NO + to give disulfide and NO via thionitrite, and (4) the NO captures molecular oxygen from air to regenerate NO2.  相似文献   

10.
The hydrothermal synthesis of V2O5, AgNO3, pyridine-2,6-dicarboxylic acid (H2pdc) and 2,2′-bipyridine (bpy) in water at 160 °C for 4 days yields a novel 1D coordination polymer VO2(C7H3O4N)Ag(C10H8N2)·H2O (1). Each V center chelates to a tridentate ligand pdc2? and coordinates to two O atoms, while the square based pyramid conformation of Ag center consists of three O atoms and a bpy molecular. V and Ag polyhedra alternate by either carboxyl or oxo bridges to further form a unique 3d–4d heterometal-based 1D double-chain ribbon.  相似文献   

11.
On the basis of quantum‐chemical calculations, it was shown that among six types of active centers (ACs) that can form during the polymerization of butadiene with lanthanide‐based catalytic systems, five types (containing electron‐accepting chlorine atoms in the coordination sphere of a lanthanide) exhibit a π‐allyl binding of the terminal unit of a growing polymer chain to a lanthanide atom and function as cis‐regulating. The sixth type of ACs is characterized by a σ‐alkyl structure and shows a trans‐stereospecificity. This results was used to interpret the data on the microstructure of polybutadiene prepared using NdCl3 · 3TBP‐Al(iso‐C4H9)3, NdCl3 · 3TBP–Mg(n‐C4H9)(iso‐C8H17) catalytic systems and their combinations (TBP is tributyl phosphate). © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 601–603, 2003  相似文献   

12.
Deactivation of Co–Ru/γ‐Al2O3 Fischer–Tropsch (FT) synthesis catalyst along the catalytic bed over 850 h of time‐on‐stream (TOS) was investigated. Catalytic bed was divided into four parts and structural changes of the spent catalysts collected from each catalytic bed after FT synthesis were studied using BET, ICP, XRD, TPR, carbon determination, H2 chemisorption and oxygen titration techniques. Rapid deactivation was observed during first 200 h of FT synthesis. In this case, the deactivation rate was not dependent on the number of the catalyst active sites. It was zero order to CO conversion and independent of the size of active sites. Beyond the TOS of 200 h, the deactivation could be simulated with a power law expression: . The physical properties of the catalyst charged in 1st half of the reactor did not change significantly. Interaction of cobalt with alumina and formation of mixed oxides of the form xCoO·yAl2O3 and CoAl2O4 was increased along the catalytic bed. Percentage reducibility and dispersion decreased by 2.4–25.5% and 0.5–8.8% for the catalyst in the beds 1 and 4, respectively. Particle diameter increased by 0.8–6.1% for the catalyst in the beds 1 and 4 respectively suggesting higher rate of sintering at last catalytic bed. The amount of coke formation in the 4th catalytic bed was 6 times more than that of in bed 1.  相似文献   

13.
In the presence of methylaluminoxane (MAO), ethylene polymerization was successfully performed with homobinuclear zirconocene complexes {[(C5H5)ZrCl2](C5H4CH2 C6H4CH2C5H4)[(C5H5)ZrCl2]; 3o , 4m , and 5p }, which were prepared conveniently by the reaction of disodium(phenylenedimethylene)dicyclopentadienide [C6H4(CH2C5H4Na)2] with 2 equiv of (N5‐Cyclopentadienyl)trichlorozirconium dimethoxyethane (CpZrCl3(DME)) in tetrahydrofuran and characterized by 1H‐NMR and elemental analysis. The effects of the polymerization parameters, such as the temperature, time, concentration of the catalyst, MAO/catalyst molar ratio, and isomeric difference of the homobinuclear metallocene complexes 3o , 4m , and 5p were studied in detail. The results showed that all three catalytic systems had moderate activities in ethylene polymerization and afforded polyethylene with relatively broad polydispersities. The catalytic activity of 4m was somewhat higher than that of 3o and 5p but lower than that of 4,4′‐bis(methylene)biphenylene‐bridged zirconocene catalysts; this indicated that the distance between the two metal centers was too short in comparison with a 4,4′‐bis(methylene)biphenylene bridge to increase the catalytic activity. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci, 2006  相似文献   

14.
As a result of Np(V) glycolate reaction with 2,2′-bipyridyl, the novel coordination polymer [NpO2(C2O3H3)(Bipy)] · 2.5H2O was formed. The structure consists of infinite chains of [NpO2(C2O3H3)(Bipy)] and water molecules. The crystal structure is confirmed by IR and UV–vis spectroscopic data. This complex is the first example of An glycolate with neutral molecular ligand.  相似文献   

15.
Summary (C5H5)2Zr(O2C)CH3 and (C5H5)2Zr(O2C)CH2CH3 complexes were synthesized, characterized and activated with MAO for ethylene polymerization. The highest catalytic activity was achieved at Al/Zr molar ratio of 3000 for both systems. The effects of the size of the R group in the carboxylate ligands, the Al/Zr molar ratio and reaction temperature on the catalytic activity and polymer properties were studied and discussed.  相似文献   

16.
Continuing our work on the synthesis of MoO2L2 and MoO3LALB that show excellent anti-cancer activities in vitro, the MoL3 have been synthesized by the solvothermal reaction of Na2MoO4 with catechols and 1,2-DPA in the mixed solvent of MeCN/MeOH. X-ray diffraction revealed that Mo in chiral octahedral geometry coordinate with three catechol ligands formed three five-membered rings, and the [Mo(C6H4O2)3] are linked by hydrogen bonds Mo(OC6H4)O…H–N(C4H8O)N–H…O(C6H4O)Mo through the by-product (C4H8N2O) that are formed by one 1,2-DPA with one CO2 on the catalysis of Mo-complex. Also, we have disassembled bulk crystal into nano-aggregates, and under TEM mono-lamella morphology of nanostructures was observed, which agrees well with the previous conclusion that the morphologies of the nano-aggregates are associated with the quantum motifs in their crystal lattices. [Mo(C6H4O2)3] have also been characterized by UV–vis spectra, cyclic voltammogram and thermogravimetric analysis.  相似文献   

17.
The shape-defined Fe3O4 nanocatalysts such as spheres and polyhedrons prepared by a simple solvothermal method without calcination were applied in Fischer–Tropsch synthesis (FTS), which showed excellent catalytic activity and C5+ selectivity compared to the traditional Fe catalyst. Especially, the Fe3O4 nanocatalyst with nanospheres (FNS) displayed higher catalytic activity and C5+ selectivity (> 64%) than the Fe3O4 nanopolyhedrons (FNP). It was found that FNS was more favorable to the reduction and dispersion of iron species as well as formation of surface carbonaceous species (especially for χ-Fe5C2) compared to FNP, which provided more active sites for FTS and facilitated the product distribution shifting towards heavy hydrocarbons.  相似文献   

18.
The effects of Co loading and calcination temperatures on the catalytic activity of Co/Al2O3 for selective catalytic reduction (SCR) of NO with ethylene in excess oxygen were investigated. Co/Al2O3 showed high and low activities when calcined at high (800 °C) and low (350 °C) temperatures, respectively. The formation and dispersion of cobalt species for catalysts calcined at 350 and 800 °C as well as for Al2O3 were studied by XRD, UV–vis and FTIR spectra. Combined with DRIFTS results of ad-species and reaction experiments, it allowed us to correlate the catalytic activity with active sites of Co/Al2O3, and the catalytic functions of active cobalt species and support were clarified. Co3O4 species contributed to the oxidation of NO to various nitrates and of C2H4 to reactive formate species, even in the absence of O2, whereas the side reaction of ethylene combustion occurred simultaneously when excess oxygen was present. Tetrahedral Co2+ ions in CoAl2O4, which acted as the active sites, were responsible for the reaction between formate and nitrate species to form organic nitro compound.  相似文献   

19.
Using the transient hot‐wire method, the thermal conductivity properties of the molten Li2O–B2O3 and K2O–B2O3 systems were measured. The thermal conductivity increases with decreasing the temperature due to the borate structure change. In addition, calculations of the one‐dimensional Debye temperature and the phonon mean free paths as a function of temperature of the alkali borate systems were made. At a fixed temperature of 1273 K, the effect of the alkali oxide concentration on the thermal conductivity was evaluated. Within a range of 10–30 mol% Li2O (or K2O), a positive relationship between the thermal conductivity and 4‐coordinate boron was obtained. However, below 10 mol% Li2O (or K2O), the change in the intermediate‐range order of the borate structure had a more dominant effect on the thermal conductivity. Finally, the effect of cations on the thermal conductivity in the various molten R2O–B2O3 (R=Li, Na and K) systems was considered. Depending on the type of cation, the change in the ionization potential had an effect on the thermal conductivity and also resulted in a change in the bond strength.  相似文献   

20.
A new model for mixtures (two and more solutes) of aqueous electrolyte solutions was found to be as accurate as other models, or more accurate, in prediction of new experimental results of the ternary systems HClO4–NaClO3–H2O and HClO4–NaCl–H2O. The water activity values are then used to study the mechanism of the chloride–chlorate reaction, generating chlorine dioxide: 2H+ + ClO?3 + Cl? → ClO2 + 1/2Cl2 + H2O Spectrophotometric measurements of the production rate of ClO2 have confirmed that the intermediary species in the proposed equilibrium H+(mH2O) → H+(m-n)H2O + nH2O Is actually H(H20)+m-n. The final kinetic expression for the reaction rate of chlorine dioxide generation, used in pulp bleaching, is then derived to explain the high order with respect to acid concentration.  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