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 93 毫秒
1.
Summary The anionic polymerization of pentafluorophenylmethyl methacrylate (PFPMA) were examined with several anionic initiators. The initiators of relatively low activity such as t-C4H9OK, organozinc compounds and Al(C2H5)3 produced poly(PFPMA) in fairly high yields though no polymer was obtained by the initiations of n-C4H9Li and C2H5MgBr within 7 days. The anionic polymerization reactivity of PFPMA was found to be strongly affected by the pentafluorophenyl group though the e-value of PFPMA is similar to those of methyl methacrylate and benzyl methacrylate.Part 7, Polym. J. 19, 958 (1987)  相似文献   

2.
Lewis hard acid AlCl3 was softened by some butyl halides forming highly polarized liquids. These liquids are similar to the ionic liquids (ILs) with metallic complex anion and varying composition, and thus termed here as carbonium pseudo ILs (CPILs). The CPILs, that is, t‐C4H9Cl‐AlCl3, n‐C4H9Cl‐AlCl3, and t‐C4H9Br‐AlCl3, show very strong desulfurization activity for various thiophenic compounds like 3‐methylthiophene, benzothiophene, and dibenzothiophene. The above thiophenic compounds can be removed completely from model oils within 20 min by a very small amount of CPILs reactive extractant. The extractive mechanism is deemed as an acid–base complexation along with alkylation of the thiophenic compounds, and the Lewis acidity comes from both carbonium ion (borderline acid) and the dissolved AlCl3 (hard acid). The t‐C4H9Cl‐AlCl3 shows good selectivity for three thiophenic compounds even in toluene‐containing system. Further, some CPILs show satisfactory desulfurization performance for the model gasoline that mimics the composition of real one. © 2012 American Institute of Chemical Engineers AIChE J, 59: 948–958, 2013  相似文献   

3.
Summary Several PMMA samples prepared in toluene with anionic initiators such as t-C4H9MgBr, n-C4H9MgCl, 1,1-diphenylhexyllithium, t-C4H9Li and t-C4H9Li(n–C4H9)3Al complex were analyzed by the on-line GPC/NMR method using a 500 MHz 1H NMR spectrometer as a real-time detector of GPC. The molecular method for a short time (60 min per sample) with a small amount of the sample (1 mg). The plots of intensities of the -methyl proton resonances due to mm- and rr-triads against elution time showed the variation of tacticity with molecular weight of the PMMA. On the basis of the results, the natures of active species in the polymerizations were discussed.  相似文献   

4.
Poly(L-lactide) (PLLA) oligo-esters with α-hydroxyl-ω-alkyl (alkyl = −CH2−[CH2−CH2]m−CH3, where m = 1, 2, 4, 5, 6, 7, 8, 9, and 10) end groups were synthesized by ring-opening polymerization of L-lactide (L-LA) catalyzed by tin(II) 2-ethylhexanoate Sn(Oct)2 in the presence of aliphatic alcohols as initiators (HO−CH2−[CH2−CH2]m−CH3, where m = 1, 2, 4, 5, 6, 7, 8, 9, and 10). High yields (~ 62 to 71%) and M n(NMR) in the range of 2120–2450 Da (PLLA) were obtained. Effects of alkyl end groups on thermal properties of the oligo-esters were analyzed by DSC, TGA and SAXS. Glass transition temperature (T g) gradually decreases with increase in the percent of−CH2−[CH2−CH2]m−CH3 end group, as results alkyl end group provides most flexibility to PLLA. An important effect of alkyl end group on a double cold crystallization (T c1 and T c2) was observed, and is directly related with the segregation phase between alkyl end group and PLLA. TGA analysis revealed that PLLA oligo-esters are more thermally stable with docosyl (−C22H45) respect to the butyl (−C4H9) end group, probably is due to steric hindrance of the end group (docosyl respect to butyl) toward intermolecular and intramolecular transesterification. SAXS analysis showed that alkyl end group as docosyl restricted the growth of lamellae thickness (D) due to steric hindrance. Characterization of hydroxyl and alkyl end groups in the PLLA oligo-esters was determined by MALDI-TOF, GPC, FT-IR and 1 H and 13 C NMR.  相似文献   

5.
The reactivity of cyclosiloxane with 3,3,4,4,5,5,6,6,6‐nonafluorohexyl group and its application to fluorosilicone synthesis were studied. In contrast to the polymerization of the commercially available 1,3,5‐tris(3,3,3‐trifluoropropyl)‐1,3,5‐trimethylcyclotrisiloxane (CF3–D3), the polymerization of 1,3,5‐tris(3,3,4,4,5,5,6,6,6‐nonafluorohexyl)‐1,3,5‐trimethylcyclotrisiloxane (C4F9–D3) with sodium hydroxide yielded 1,3,5,7‐tetrakis(3,3,4,4,5,5,6,6,6‐nonafluorohexyl)‐1,3,5,7‐tetramethylcyclotetrasiloxane {[C4F9CH2CH2(CH3)SiO]4 (C4F9–D4)} as the major product, which was the result of the depolymerization by a backbiting mechanism. On the other hand, the polymerization of C4F9–D3 with trifluoromethanesulfonic acid yielded a polymer with good reproducibility, although the distribution of molecular weight tended to be bimodal. C4F9–D3 was also successfully copolymerized with octamethylcyclotetrasiloxane (D4). The properties of the obtained copolymers changed almost linearly with the amount of [C4F9C2H4(CH3)SiO] units, except for the surface tension. The surface tension of the copolymer containing more than 50% [C4F9C2H4(CH3)SiO] units by weight was almost as low as that of the homopolymer. When the copolymer containing 50% [C4F9C2H4(CH3)SiO] units by weight was examined for an antifoaming agent, it worked as effectively as the homopolymer in the toluene foaming system. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 82: 3333–3340, 2001  相似文献   

6.
Summary Polymerization of MMA with t-C4H9Li in toluene in the presence of trialkylaluminum(alkyl=methyl, ethyl, butyl, isobutyl, and octyl) was examined at various Al/Li ratios. Triethyl-, tributyl and trioctylaluminums worked quite effectively in the polymerization to give highly syndiotactic PMMA with narrow molecular weight distribution (MWD) at the Al/Li ratio of 3 and higher than 3. The highest syndiotacticity of 96% was attained with t-C4H9Li/(n-C8H17)3Al (1/3) at –93°C. The polymerization was initiated by t-C4H9 anion and proceeded in a living manner. The structure of the syndiotactic PMMA is chemically identical to that of isotactic PMMA prepared with t-C4H9MgBr. Consequently, highly isotactic and syndiotactic PMMAs with narrow MWD and with the same chemical structure from -end to -end were obtained.  相似文献   

7.
Co–BaCO3 catalysts exhibited high catalytic performance for oxidative dehydrogenation of ethane (ODE) using CO2 as oxidant. The maximal formation rate of C2H4 was 0.264 mmol · min−1 · (g · cat.)−1 (48.0% C2H6 conversion, 92.2% C2H4 selectivity, 44.3% C2H4 yield) on 7 wt% Co–BaCO3 catalyst at 650 °C and 6000 ml. (g · cat.)−1. h−1. Co–BaCO3 catalysts were comparatively characterized by XRF, N2 isotherm adsorption-desorption, XRD, H2-TPR and LRs. It was found that Co4+–O species were active sites on these catalysts in ODE with CO2. The redox cycle of Co–O species played an important role on the catalytic performance of Co–BaCO3 catalysts. On the other hand, the co-operation of BaCO3 and BaCoO3 was considered to be one of possible reasons for the high catalytic activity of these catalysts.  相似文献   

8.
Preparation of poly(butyl acrylate)‐grafted oil palm empty fruit bunch fibre (OPEFB) has been successfully carried out using H2O2/Fe2+ as a redox initiator in aqueous solution. The effects of reaction temperature and period, as well as the amount of monomer, initiator and nitric acid, on the percentage of grafting were investigated. The percentage of grafting increases with amount of monomer and can be controlled by setting the appropriate reaction conditions. The maximum percentage of grafting (about 265%) was obtained when the reaction was carried out without acid under optimum conditions (reaction period: 2 h, reaction temperature: 55 °C, amount of H2O2: 5.88 mmol) with 42.2 mmol of monomer. Mechanisms of grafting of butyl acrylate onto OPEFB were proposed. Characterization of the grafted OPEFB was done by Fourier‐transform infrared spectroscopy and scanning electron microscopy. The thermal properties were studied by thermogravimetric analysis. Copyright © 2003 Society of Chemical Industry  相似文献   

9.
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   

10.
Sulfonated thiophenes, sodium 2-(3-thienyloxy)ethanesulfonate (C6H7S2O4Na) and sodium 6-(3-thienyloxy)hexanesulfonate (C10H15S2O4Na), were synthesized and used in the fabrication of ion-selective electrodes (ISEs) sensitive and selective to Ag+. The Ag+-ISEs were prepared by galvanostatic electropolymerization of 3,4-ethylenedioxythiophene (EDOT) on glassy carbon (GC) electrodes, with either C6H7S2O4 or C10H15S2O4 as the charge compensator (doping ion) for p-doped poly(3,4-ethylenedioxythiophene) (PEDOT). Potentiometric measurements were carried out with these sensors, GC/PEDOT(C6H7S2O4) and GC/PEDOT(C10H15S2O4), to study and compare their sensitivity and selectivity to silver ions. PEDOT(C6H7S2O4) and PEDOT(C10H15S2O4) films were also studied by using other techniques such as cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), electrochemical quartz crystal microbalance (EQCM) and Fourier transform infrared spectroscopy (FTIR).Results from the potentiometric measurements showed that the difference in length of the alkyl chain of the doping ions C6H7S2O4 and C10H15S2O4 has no significant effect on the sensitivity or selectivity of GC/PEDOT(C6H7S2O4) and GC/PEDOT(C10H15S2O4) sensors to Ag+. More differences can be seen in the cyclic voltammograms and EIS spectra of the sensors. FTIR spectra confirmed that both C6H7S2O4 and C10H15S2O4 act as doping ions in the electrosynthesis of PEDOT-based films and they are not irreversibly immobilized in the polymer backbone.  相似文献   

11.
The complete photocatalytic oxidation of C2H4 with O2 into CO2 and H2O has been achieved on ultrafine powdered TiO2 photocatalysts and the addition of H2O was found to enhance the reaction. The photocatalytic reaction has been studied by IR, ESR, and analysis of the reaction products. UV irradiation of the photocatalysts at 275 K led to the photocatalytic oxidation of C2H4 with O2 into CO2, CO, and H2O. The large surface area of the photocatalyst is one of the most important factors in achieving a high efficiency in the photocatalytic oxidation of C2H4. The photoformed OH species as well as O 2 and O 3 anion radicals play a significant role as a key active species in the complete photocatalytic oxidation of C2H4 with O2 into CO2 and H2O. Interestingly, small amount of Pt addition to the TiO2 photocatalyst increased the amount of selective formation of CO2 which was the oxidation product of C2H4 and O2.  相似文献   

12.
Summary Polymerization of methyl methacrylate (MMA) withtert-butyllithium (t-C4H9Li) in toluene in the presence of aluminum alkoxides such as ethoxide,tert-butoxide and 2,6-di-tert-butylphenoxide, were examined at various Al/Li ratios. In the cases of ethoxide andtert-butoxide, predominantly isotactic polymers with broad molecular weight distribution were obtained. Combinations oft-C4H9Li and bis(2,6-ditert-butylphenoxy)methylaluminum [MeAl(ODBP)2] were found to be an efficient initiating system for heterotactic polymerization of MMA, which gives PMMA rich in heterotactic triads up to 68% with narrow molecular weight distribution (Mw/Mn=1.09–1.17). End group analysis by1H NMR indicated thatt-C4H9Li initiates the polymerization and MeAl(ODBP)2 works as a stereospecific modifier. From stereosequence analysis of the heterotactic PMMA by13C NMR, it was found that the calculated pentad fractions from the first-order Markovian statistics (Pm/r=0.742, Pr/m=0.627) fitted the observed ones better than those from Bernoullian statistics. The glass transition temperature of the heterotactic PMMA was 13°C lower than that of syndiotactic PMMA, and the intrinsic viscosity in tetrahydrofuran was close to that of isotactic PMMA with a similar molecular weight but higher than that of syndiotactic PMMA.  相似文献   

13.
A series of organotin(IV) carboxylates complexes; namely, [(Me2Sn)4O2(RCOO)4] (R = C12H15 1, C9H11 2, C8H8ClO 3, C7H9 4) and [Me3(RCOO)]n (R = C12H15 5, C9H11 6, C8H8ClO 7, C7H9 8) have been synthesized. All complexes were characterized by elemental analysis, FT-IR, and NMR (1H, 13C and 119Sn) spectroscopy. Among them, the structures of complexes 13 and 58 were also determined by X-ray crystallography. The structural analysis showed that complexes 13 are the same tetranuclear monomer, and complexes 58 are the same 1D zigzag chain coordination polymer. Furthermore, each complex 1, 2 and 3, can form a supramolecular chain through weak intermolecular interactions.  相似文献   

14.
The reaction of (NO + C3H8 + O2) can result in selective formation of NO2 over H-ZSM5, Cu,H-ZSM5, Ag,H-ZSM5, and Li,H-ZSM5 catalysts when the concentrations of NO and O2 are 0.1 and 9%, SV > 60,000 h−1 (typical for automotive exhausts), and C3H8/NO > 1. Despite stoichiometric excess of reductant hydrocarbon below this limit, the in situ formed NO2 does not react with C3H8, thus conversion of NO to N2 is negligible. NO can be reduced by C3H8 selectively to N2 only when C3H8/NO ≧ 1. Contrary to many suggestions the reaction temperature, concentration of oxygen, space velocity, and type of exchange ions have minor influence on the selectivity for N2. These parameters affect the rates of reactions (NO + 2), (C3H8 + NOx) and (C3H8 + O2), therefore they also affect the production of N2 in the HC-SCR process, but only when the ratio of C3H8/NO permits. The metal-exchanged zeolites were prepared in situ by solid-state ion exchange from H-ZSM5. Despite the low degree of copper exchange (63%), Cu,H-ZSM5 produces substantially more N2 than H-ZSM5, Ag,H-ZSM5, or Li,H-ZSM5. However, the selectivity for N2 is lowest over Cu,H-ZSM5, which also produces considerable NO2 in the reaction of (NO + C3H8 + O2) even at C3H8/NO ≧ 1. Contrary to prior findings, the catalytic activity of Cu,H-ZSM5 for the oxidation of NO by O2 to NO2 in absence of hydrocarbon was comparable to that of H-ZSM5 at high space velocities (2.3 l g−1 min−1). By replacing 30 and 40% of the protons of H-ZSM5 by Ag+ and Li+ ions in Ag,H-ZSM5 and Li,H-ZSM5, respectively, the catalytic activity for this reaction becomes negligible at temperatures ≧100°C. Some mechanistic consequences of these experimental observations are discussed. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

15.
The oxidative properties of Li/MgO in the absence of O2 were studied at 730° C using C2H4 as a reducing agent and a multisectional flow stainless steel tubular reactor. Large amounts of CO and H2 were determined. It was demonstrated that Li/MgO exhibited a high oxygen mobility. The exchange capacity was (2.4–4.0) × 1020 oxygen atoms per 1 g of the catalyst and the period of oxygen donation was 150 h or more. After the catalyst reduction, its oxygen transfer ability was fully restored by re-oxidation in the stream of air. Oxidation of C2H4 to CO and H2 was accompanied by its decomposition to C and H2. The ratio H2/C2H4 was found to be 1.91 ± 0.13 independently of the oxidation state of the catalyst, the location of the sampling point along the catalyst bed (and the post-catalytic zone) and the duration of the experiment. The mechanism was discussed.  相似文献   

16.
6-dimethylamino-6-methylfulvene (7) was converted to the [(C5H4)–CMe2–NMe2] ligand system (8) by treatment with methyllithium. Its reaction with MCl4 (M = Zr, Ti) followed by treatment with CH3Li gave the respective [(C5H4)–CMe2–NMe2]2M(CH3)2 complexes (12). Their reaction with B(C6F5)3 led to reactive metallocene cation complexes that instantaneously underwent CH activation at a N–CH3 group to yield the metallacyclic cation complexes 15. (tert-butylaminomethyl)fluorene was prepared by the addition of tert-butylisocyanate to fluorenyllithium followed by hydride reduction. Deprotonation by a variety of bases gave rise to a series of competing and consecutive reactions to yield several unusually structured products, among them a fluorenyl-anellated η5-1-azapentadienyl anion equivalent (25) and [(flu)-CH2–NCMe3]Li2 (23). An improved way of generating synthetically useful C1-linked [Cp–C1(R) n –NR1]2- dianion equivalents was developed starting from 6-amino-6-methylfulvene (26). N-silylation followed by double deprotonation with, e.g., lithium diisopropylamide cleanly furnished the respective [(C5H4)–C(=CH2)–NSiMe3]2- dianion 33 (isolated as the dilithio derivative). Its reaction with Cl2Zr(NEt2)2 in THF gave [η5:κ-N-(C5H4)–C(=CH2)–NSiMe3]Zr(NEt2)2 36. Activation of 36 with methylalumoxane in toluene led to the formation of a C1-linked “constrained geometry” Ziegler catalyst that polymerized ethylene similarly as the [(C5Me4)SiMe2NCMe3]ZrCl2 derived literature system. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

17.
Dehydrogenation of isobutane (iC4) to isobutene has recently received considerable attention because of the increasing demand for methyl tertiary butyl ether (MTBE) and ethyl tertiary butyl ether (ETBE), as additives for gasoline to increase the octane number and to substitute lead. Among several catalysts, Pt–Sn/γ-Al2O3 catalyst is one of the most suitable ones. In this study, Pt–Sn, Pt–Sn–K and Pt–Sn–Li catalysts supported on γ-Al2O3 have been investigated. The catalysts were prepared by co-impregnation using H2PtCl6, SnCl2 as the metal precursors and K2CrO4 and LiOH as the promoters. The obtained catalysts were characterized by TPR and XRD techniques. Reaction temperature, feed flow rate (Q), iC4/H2 ratio and the catalyst type were the main variables that were investigated. Initially, primary experiments were carried out for the identification of the optimum range of operating variables, then, Taguchi's algorithm was employed to design the experiment. The results revealed that the Pt–Sn–K catalyst showed higher selectivity than Pt–Sn–Li. Under the optimum operating conditions i.e. at 550 °C, feed flow rate of 50 cm3/min and iC4/H2 ratio of 1/3, selectivity toward isobutene of higher than 90% was achieved with the corresponding conversion values of about 15%.  相似文献   

18.
Vapor of propargyl alcohol under reduced pressure was glow-discharged for plasma polymerization by which hydrophilic polymer film could be deposited onto a solid substrate. The hydrophilic property was more enhanced with use of the same monomer containing water. An interpretation of chemical structures of the polymer molecules was therefore attempted to discover the reason of the hydrophilic property. The polymer materials were subjected to IR spectroscopy, 13C-NMR and 1H-NMR spectroscopies, elemental analysis, number average molecular weight determination, mass spectrometry, and other wet chemical analysis for . The IR spectra of the polymers showed the presence of OH, C?O, and C?C groups with a disappearance of C?C group of the monomer. It was also found that the molecular formula of a solvent extract of the polymer derived from the plain monomer was C26H36O9 ≒ 490, involving four hydroxyl, one carbonyl, and two to three olefinic unsaturation groups within a unit molecule. On the other hand, the polymer from 60% water-containing monomer showed its molecular formula as C44H66O18 ≒ 890, containing seven to eight hydroxyl, two carbonyl, and one to two olefinic unsaturation groups. This information was used to establish the symbolic chemical structures of the polymer molecules. It has been found that both polymers were not as much crosslinked as other plasma polymers because of having side chains with fewer branching. The oxygen atoms were retained in the polymer not only as the forms of OH and C?O but also as a member of the principal frame of the polymers.  相似文献   

19.
A novel bithiazole oligomer (PCBT) was synthesized from C60 and the diazo salt of 2,2′-diamino-4,4′-bithiazole (DABT). Its ferro-complex (PCBT-Fe2+) was prepared from PCBT and FeSO4 in DMSO solution under a purified nitrogen atmosphere. The magnetic behavior of PCBT and PCBT-Fe2+ was measured as a function of magnetic field strength (0-60 kOe) at 5 K and as a function of temperature (5-300 K) at a magnetic field strength of 30 kOe. PCBT-Fe2+ complex exhibits a hysteresis cycle at 5 K, the observed coercive field (HC) and remnant magnetization (Mr) are 690 Oe and 0.12 emu/g, respectively. The results show that PCBT is an anti-ferromagnet and its Fe2+-complex is a soft ferromagnet.  相似文献   

20.
The effect of Cl? ion addition to a Li+-ZnO catalyst has been studied with respect to the oxidative coupling of CH4 and the oxidative dehydrogenation (OXD) of C2H6. Increasing the Cl/Li ratio from 0.65 to 0.90 had relatively little effect on the CH4 conversion, whereas the C2H4/C2H6 ratio was enhanced significantly as a result of an increase in the OXD reaction rate. Conversely, loss of Cl? from the catalyst during the reaction had a much more deleterious effect on ethane OXD activity than on methane coupling activity. Addition of Cl? ions at a Cl/Li ratio of 0.9 caused a decrease both in the number of basic sites and in the basic strength of these sites, as determined by temperature-programmed desorption of CO2. The similarities between the results obtained over Li+-ZnO-Cl? catalysts and those previously reported for Li+-MgO-Cl? catalysts confirm that basicity of the host oxide plays only a minor role in determining the properties of these chlorided catalysts.  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