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.
In this work, the reversible addition-fragmentation chain transfer (RAFT) polymerization of vinyl acetate (VAc) was successfully performed at room temperature using 60Co γ-irradiation as the initiation source. Under the dose rate of 10 Gy/min irradiation, the polymerization proceeded smoothly and converted approximately 90% of the monomer within 7 h. The molecular weight distribution (Mw/Mn) remained narrow (Mw/Mn < 1.35) up to 90% conversion. Compared to AIBN-initiated RAFT polymerization at 60 °C, 60Co γ-irradiation-initiated RAFT polymerization is a technique that can better control the molecular weight, especially at high conversion. The 1H NMR spectra and matrix-assisted laser desorption/ionization time-of-flight mass spectrometry confirmed that most of the chain ends of poly(VAc) (PVAc) from γ-irradiated RAFT polymerization were living and can be reactivated for chain-extension reactions. The microstructures of PVAc from 60Co γ-irradiated RAFT polymerization (almost head-to-tail addition) and AIBN-initiated RAFT polymerization (5% tail-to-tail addition) were different, as revealed by the 13C NMR spectra. For the first time, 60Co γ-irradiation was used as an initiation source for RAFT polymerization of VAc at room temperature.  相似文献   

2.
(S)-2-(Ethyl propionate)-(O-ethyl xanthate) (X1) and (S)-2-(Ethyl isobutyrate)-(O-ethyl xanthate) (X2) were used as the reversible addition-fragmentation chain transfer (RAFT) agents for the radical polymerization of vinyl acetate (VAc). The former showed the better chain transfer ability in the polymerization at 60°C. Kinetic study with both RAFT agents showed pseudo-first order kinetics up to around 85% monomer conversion. Molecular weight of the resulting polymer increased linearly with increase in the monomer conversion up to around 85%. The observed molecular weights calculated from 1H-NMR spectrum [Mn(NMR)] are close to the corresponding theoretical molecular weights [Mn(theor)]. The corresponding polydispersity index (PDI) of the resulting polymers remained almost constant at around 1.2 up to ∼ 65% monomer conversion and then increased gradually with the further increase in the monomer conversion. Chain-end analysis of the resulting polymers by 1H-NMR showed clearly that polymerization started with the radical forming out of the xanthate mediator. The negligible homo-chain extension and the hetero-chain extension involving synthesis of poly(VAc)-b-poly(NVP) diblock copolymer were occurred. © 2012 Wiley Periodicals, Inc. J Appl Polym Sci, 2012  相似文献   

3.
No alternating copolymers of vinyl acetate (VAc) and butyl acrylate (BA) were obtained by free radical copolymerization in the presence of GeCl4 and BCl3 (compared with the acrylic acid–vinyl acetate copolymerization system). By ultraviolet spectral analysis, it was concluded that both BCl3 and GeCl4 can form complexes with butyl acrylate. The BA–BCl3 complex constants were determined by 1H NMR; KB=33·2 (25°C). The reason for the gel formation in the BA–Vac–BCl3 copolymerization system was discussed. When vinyl acetate reacted with BCl3, cationic polymerization probably occurred. A white gel product probably resulted from the polymerization of the BA–BCl3 complex. © 1998 SCI.  相似文献   

4.
Homopolymers of N-vinylpyrrolidone (VP) and vinyl acetate (VAc) were synthesized by a free-radical solution polymerization technique. Copolymers of VP and VAc in various monomer feed ratios were also synthesized by the same procedure. They were characterized by elemental analysis, FTIR, PNMR, TGA, swelling, and viscosity measurements. The reactivity ratios of the monomers were computed by both Fineman–Ross and Kelen–Tudos methods using data from both PNMR and elemental analysis studies. The activation energy values for various stages of decomposition were calculated from TGA analysis using Broido's method. The viscosity measurements were carried out at four different temperatures: 30, 35, 40, and 45°C. The activation parameters of the viscous flow, voluminosity (VE), and shape factor (ν) were also computed for all systems. © 1998 John Wiley & Sons, Inc. J Appl Polym Sci 68:91–102, 1998  相似文献   

5.
The attachment of anticancer agents to polymers is a promising approach towards reducing the toxic side‐effects and retaining the potent antitumour activity of these agents. A new tetrahydrophthalimido monomer containing 5‐fluorouracil (ETPFU) and its homopolymer and copolymers with acrylic acid (AA) and with vinyl acetate (VAc) have been synthesized and spectroscopically characterized. The ETPFU contents in poly(ETPFU‐co‐AA) and poly(ETPFU‐co‐VAc) obtained by elemental analysis were 21 mol% and 20 mol%, respectively. The average molecular weights of the polymers determined by gel permeation chromatography were as follows: Mn = 8900 g mol?1, Mw = 13 300 g mol?1, Mw/Mn = 1.5 for poly(ETPFU); Mn = 13 500 g mol?1, Mw = 16 600 g mol?1, Mw/Mn = 1.2 for poly(ETPFU‐co‐AA); Mn = 8300 g mol?1, Mw = 11 600 g mol?1, Mw/Mn = 1.4 poly(ETPFU‐co‐VAc). The in vitro cytotoxicity of the compounds against FM3A and U937 cancer cell lines increased in the following order: ETPFU > 5‐FU > poly(ETPFU) > poly(ETPFU‐co‐AA) > poly(ETPFU‐co‐VAc). The in vivo antitumour activities of all the polymers in Balb/C mice bearing the sarcoma 180 tumour cell line were greater than those of 5‐FU and monomer at the highest dose (800 mg kg?1). © 2002 Society of Chemical Industry  相似文献   

6.
Bimetallic salicylaldimine-nickel complexes, 2,4,6-Me3-1,3-{[NCH–(3′-R-5′-Y-2′-O–C6H3)-κ2-N,O]Ni(Ph) (PPh3)}2 [R = tert-Bu, Y = Me, 1b; R = Ph, Y = H, 2b] were prepared and their catalytic behaviors of ethylene polymerization were investigated. The bimetallic complex 2b shows higher activities (2.9 × 105 g PE mol−1 Ni h−1) for ethylene polymerization and affords polymer with high molecular weight (Mw = 1.41 × 105) and broad molecular weight distribution (Mw/Mn = 6.1) than its mononuclear matrix, {[(2,6-Me2C6H3)–NCH–(3′-Ph-2′-O–C6H3)-κ2-N,O]Ni(Ph)(PPh3)} (3) (Activity = 5.5 × 104 g PE mol−1 Ni h−1; Mw = 1.86 × 104; Mw/Mn = 2.8).  相似文献   

7.
The bulk copolymerizations of acrylonitrile (AN) with vinyl acetate (VAc) initiated by azobisisobutyronitrile (AIBN) and the suspension copolymerization of AN with VAc and sodium methallylsulfonate (SMAS) in water with a Na2S2O5–Na2ClO3 redox initiator system at 65°C, were investigated. The copolymer compositions were determined by 1H-NMR. The reactivity ratios (γs) for the two copolymerization systems were determined analytically, based on Mayo–Lewis equation, by fitting the calculated curves with the experimental data. The γs for the AN and VAc bulk copolymerization were found to be γ12 = 2.85 and γ21 = 0.11. The values of the apparent γs for the suspension copolymerization of AN, VAc, and SMAS were as follows: γ12 = 3.58, γ21 = 0.39, γ13 = 1.45, γ31 = 0, γ23 = 0.92, and the rate constant ratio R3 = k31/k32 = 0.04. A simulated result produced with the obtained γs agreed fairly well with experimental data of bulk copolymerization in a batch reactor. The apparent γs obtained were also successfully used to analyze the results of suspension polymerization in a continuous pilot reactor.© 2001 John Wiley & Sons, Inc. J Appl Polym Sci 82: 854–860, 2001  相似文献   

8.
The thermal stability and nonisothermal thermal decomposition kinetics of the vinyl chloride–vinyl acetate copolymer synthesized via microsuspension polymerization (MS–VC/VAc) were studied by dynamic TG. The results showed that MS–VC/VAc gives off hydrogen chloride and acetic acid (HAc) at the same time when it was heated. VAc content produces little effect on the thermal stability and the kinetic parameters. The activation energy for giving off HCl and HAc is in the range of 120–160 kJ/mol and the Arrhenius frequency factor 3.27 × 109–9.72 × 1012 s−1. For different heating rates, the kinetic model function of the thermal decomposition obeys the Avrami–Erofeev model equation, that is, [−ln(1 − a)]1/m for g(a), where m = 0.65–0.85. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 1057–1062, 2000  相似文献   

9.
Yoshikatsu Tsuchiya 《Polymer》2008,49(5):1180-1184
Bulk polymerization of vinyl chloride (VC) with CpTi(OPh)3/MAO catalyst was investigated. The bulk polymerization of VC with CpTi(OPh)3/MAO catalyst proceeded to give poly(vinyl chloride) (PVC) with high molecular weight in good yields. The Mn of the polymer increased in direct proportion to polymer yields and the line passed through the origin. The Mw/Mn of the polymer decreased with an increase of polymer yield. The GPC elution curves were unimodal and the whole curves shifted clearly to the higher molecular weight as a function of reaction time. This indicates that the control of molecular weight can be achieved in the polymerization of VC with CpTi(OPh)3/MAO catalyst even in bulk. The structure of PVC obtained from the bulk polymerization of VC with CpTi(OPh)3/MAO catalyst consists of a regular structure. The thermal stability of the polymer obtained with CpTi(OPh)/MAO catalyst was higher than that of PVC obtained from radical polymerization and depended on the molecular weight of the polymer. In contrast to that, the initial decomposition temperature of the polymer obtained from a radical polymerization did not depend on the molecular weight. We presumed that the decomposition of the polymer obtained with CpTi(OPh)3/MAO catalyst initiated at the chain end.  相似文献   

10.
Homo‐ and copolymers of vinyl esters including vinyl acetate (VAc) and vinyl benzoate (VBz) were synthesized via the reverse iodine transfer radical polymerization technique. Polymerization was carried out in the presence of iodine as the in situ generator of the transfer agent and 2,2′‐azobis(isobutyronitrile) as the initiator at 70 °C. Reverse iodine transfer radical homopolymerization of VAc and VBz led to conversions of 76 and 57%, number‐average molecular weights of 8266 and 9814 g mol?1 and molecular weight distributions of 1.58 and 1.49, respectively. The microstructure of the synthesized polymers was investigated in detail using gel permeation chromatography, 1H NMR, 13C NMR and distortionless enhancement of polarization transfer (135° decoupler pulse) techniques. Relatively narrow molecular weight distribution and controlled and predictable trend of molecular weight versus conversion were observed for the synthesized polymers, showing that reverse iodine transfer radical homo‐ and copolymerization of VAc and VBz proceeded with controlled characteristics. Results of molecular weight and its distribution along with the 1H NMR spectra recorded for homo‐ and copolymers indicated that side reactions can occur during the course of polymerization with a significant contribution when VAc, even in a small amount, was present in the reaction mixture. This can result in polymer chains with aldehyde dead end and broadening of the molecular weight distribution. © 2015 Society of Chemical Industry  相似文献   

11.
Summary Trimethylsilyl iodide in conjunction with zinc iodide (Me3SiI/ZnI2) as an initiating system led to living cationic polymerization of isobutyl vinyl ether in toluene at 0 or –40°C or in methylene chloride at –40°C (ZnI2 was dissolved in acetone). The number-average molecular weight of the polymers was directly proportional to monomer conversion and in excellent agreement with the calculated value assuming that one polymer chain forms per unit trimethylsilyl iodide. At room temperature (+25°C), however, the polymerization failed to give perfectly living polymers; the polymer molecular weight was smaller than the calculated value. On addition of a fresh feed of monomer at the end of the polymerization at –40°C, the added feed was smoothly polymerized at nearly the same rate as in the first stage, and the polymer molecular weight continued to increase in direct proportion to monomer conversion. Throughout the reaction, the molecular weight distribution of the polymers stayed very narrow (Mw/Mn< 1.1).Living cationic polymerization of vinyl ethers by electrophile/Lewis acid initiating systems, part 2. For part 1 see ref. 2  相似文献   

12.
A crosslinkable terpolymer P(MMA‐BA‐HEMA) was prepared by atom transfer radical copolymerization of 2‐hydroxyethyl methacrylate, methyl methacrylate and butyl acrylate. The structure of the terpolymer was characterized by 1H NMR and gel permeation chromatography. The effects on the polymerization of ligand, initiator, solvent, CuCl2 added in the initial stage and reaction temperature were investigated. The optimal reaction conditions were ethyl 2‐bromopropionate as initiator, CuCl/PMDETA as catalyst, cyclohexanone as solvent, catalyst/ligand = 1:1.5, [M]0:[I]0 = 200:1 and temperature 70 °C. The reaction followed first‐order kinetics with respect to monomer concentration, indicating the best control over the polymerization process, a constant concentration of the propagating radical during the polymerization, efficient control over Mn of the polymer and low polydispersity (Mw/Mn < 1.3). © 2013 Society of Chemical Industry  相似文献   

13.
A new monomer, exo‐3,6‐epoxy‐1,2,3,6‐tetrahydrophthalimidocaproic acid (ETCA), was prepared by reaction of maleimidocaproic acid and furan. The homopolymer of ETCA and its copolymers with acrylic acid (AA) or with vinyl acetate (VAc) were obtained by photopolymerizations using 2,2‐dimethoxy‐2‐phenylacetophenone as an initiator at 25 °C. The synthesized ETCA and its polymers were identified by FTIR, 1H NMR and 13C NMR spectroscopies. The apparent average molecular weights and polydispersity indices determined by gel permeation chromatography (GPC) were as follows: Mn = 9600 g mol?1, Mw = 9800 g mol?1, Mw/Mn = 1.1 for poly(ETCA); Mn = 14 300 g mol?1, Mw = 16 200 g mol?1, Mw/Mn = 1.2 for poly(ETCA‐co‐AA); Mn = 17 900 g mol?1, Mw = 18 300 g mol?1, Mw/Mn = 1.1 for poly(ETCA‐co‐VAc). The in vitro cytotoxicity of the synthesized compounds against mouse mammary carcinoma and human histiocytic lymphoma cancer cell lines decreased in the following order: 5‐fluorouracil (5‐FU) ≥ ETCA > polymers. The in vivo antitumour activity of the polymers against Balb/C mice bearing sarcoma 180 tumour cells was greater than that of 5‐FU at all doses tested. © 2001 Society of Chemical Industry  相似文献   

14.
The emulsion polymerizations of methacrylic acid (MAA), vinyl acetate (VAc), and methyl acrylate (MA) in different VAc/MA molar concentrations (X1, 0.232/0.813; X2, 0.348/0.697; X3, 0.456/0.581; X4, 0.581/0.465; X5, 0.697/0.348) and fixed MAA concentration (0.116 mol) were carried out using sodium dodecyl sulfate (SDS; 34 mmol) as emulsifier and potassium persulphate (K2S2O8; 37 mmol) as initiator at 70 °C for 6 h in semicontinuous reaction mode. The average molecular weights (Mn, Mw) and the molecular weights distributions were determined using gel permeation chromatography (GPC). All terpolymers prepared showed monomodal molecular weights distributions. Glass transition (Tg), crystallization (Tc), and melting (Tm) temperatures and thermal stability of the prepared terpolymers were determined using differential scanning calorimetry (DSC) and thermal gravimetric analysis (TGA), respectively. The elongation percentage at rupture, tensile strength and accelerated thermal aging were determined for X1–5 terpolymers, as functions of the molar composition in the emulsion polymerization feed. The terpolymers prepared, X1–5 terpolymers, were tested as stiffening agents for cotton woven fabrics through crosslinking to cellulose, using sodium hypophosphite monohydrate (NaH2PO2·H2O) as catalyst. The effect of X1–5 compositions upon the stiffening efficiency was discussed. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 2006  相似文献   

15.
A novel emulsion copolymer of vinyl acetate (VAc) and 1‐hexene was synthesized at ambient pressure. The feeding technique, initiation system and reaction time of the copolymerization were optimized based on molecular characteristics such as the weight contribution of 1‐hexene in the copolymer chains and glass transition temperature (Tg) as well as on bulk properties like minimum film‐formation temperature (MFFT) and solid content. According to nuclear magnetic resonance spectroscopy and differential scanning calorimetry results, the combination of starve feeding and redox initiation, within a reaction time of 4 h, effectively led to the copolymerization at ambient pressure between highly reactive polar VAc monomers and non‐polar 1‐hexene monomers of low reactivity. The copolymer showed a lower Tg and MFFT, and a reasonable solid content compared to the poly(vinyl acetate) (PVAc) homopolymer. The consumption rate, hydrolysis of acetate groups and chain transfer reactions during the polymerization were followed using infrared spectroscopy. Based on the results, the undesirable reactions between the VAc blocks were hindered by the neighbouring 1‐hexene molecules. Tensile testing revealed an improvement in the toughness and elongation at break of VAc–1‐hexene films compared to PVAc films. © 2014 Society of Chemical Industry  相似文献   

16.
Summary Well-defined living polymers of isobutyl vinyl ether were obtained in the polymerization initiated with ethylaluminum dichloride (EtAlCl2) in conjunction with a stoichiometric excess of dioxane (5–10 vol%) in n-hexane at 0°C. Under these conditions, the number-average molecular weight of the polymers increased in direct proportion to monomer conversion, while the molecular weight distribution stayed narrow (Mw/Mn = 1.1–1.25). In sharp contrast, the EtAlCl2-initiated polymerization in the absence of dioxane led to non-living polymers with a broad molecular weight distribution. It was concluded that the propagating carbocation is stabilized not by the counteranion but by an externally added basic compound (dioxane) that strongly interacts with the active end.  相似文献   

17.
The self‐condensing vinyl polymerization of styrene and an inimer formed in situ by atom transfer radical addition from divinylbenzene and 2‐bromoisobutyl‐tert‐butyrate using atom transfer radical polymerization technique was studied. To study the polymerization mechanism and achieve high molecular weight polymer in a high polymer yield, the polymerization was carried out in bulk at 80°C. Proton nuclear magnetic resonance (1H‐NMR) spectroscopy and gel permeation chromatography (GPC) coupled with multiangle laser light scattering (MALLS) were used to monitor the polymerization process and characterize the solid polymers. It is proved that the polymerization shows a “living” polymerization behavior and the crosslinking reaction has been restrained effectively due to the introduction of styrene. Polymers with high molecular weight (Mw.MALLS > 105) can be prepared in high yield (near 80%). Comparison of the apparent molecular weights measured by GPC with the absolute values measured by MALLS indicates the existence of branched structures in the prepared polymers. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci 2007  相似文献   

18.
A heterogenized Wacker catalyst system in which pores of a high surface area alumina were filled with an aqueous solution of PdCl2CuCl2 was active for the oxidation of CO near room temperature. The structure of the catalyst was studied by X-ray absorption fine structure (XAFS). The active phase of Pd was a Pd11 species containing chlorine and, probably, carbonyl ligands. Direct interaction of PdPd or PdCu was not detected. The active phase of copper was found to be solid Cu2Cl(OH)3 particles in agreement with the X-ray diffraction (XRD) results. The presence of Cu was essential to keep the Pd in the Pd11 state during the reaction. The rates of CO oxidation measured at temperatures of 30–70°C showed a minimum at 40°C, which was attributable partly to an unusual structure change of the active palladium species during the reaction at this temperature.  相似文献   

19.
The synthesis of diethyl carbonate (DEC) by the oxidative carbonylation of ethanol was investigated using catalysts prepared by the dispersion of CuCl2 and PdCl2 on amorphous carbon promoted with KCl and NaOH. Catalysts were characterized extensively by XRD, XAFS, SEM and TEM with the aim of establishing their composition and structure after preparation, pretreatment, and use. It was observed that after preparation and pretreatment in He at 423 K copper is present almost exclusively as Cu(I), most likely in the form of [CuCl2] anions, whereas palladium is present as large PdCl2 particles. Catalysts prepared exclusively with copper or palladium chloride are inactive for DEC synthesis, indicating that both components must be present together. Evidence from XANES and EXAFS suggests that the DEC synthesis may occur on [PdCl2−x][CuCl2]x species deposited on the surface of the PdCl2 particles. As-prepared catalysts exhibit an increase in DEC synthesis activity and selectivity with time on stream, but then reach a maximum activity and selectivity, followed by a slow decrease in DEC activity. The loss of DEC activity is accompanied by a loss in Cl from the catalyst and the appearance of paratacamite.  相似文献   

20.
Studies were performed of tungsten carbide hydrogen-diffusion electrodes operating as anodes in electrolytic baths for regeneration of etching solutions of CuCl2 and FeCl3. Under conditions of electrolytic regeneration of copper chloride solutions (i = 40 mA cm–2, 40° C) after 1500 h operation the electrode polarization increased by about 200 mV. Maximum current efficiency of 60–65% was obtained at I k = 80 mA cm–2. It is demonstrated that the replacement of the standard carbon anodes with tungsten carbide hydrogen-diffusion electrodes and the elimination of the ion exchange membrane separating the anodic from the cathodic space leads to a 2–4 V decrease of the electrolytic bath voltage. The regenerated solutions of CuCl2 and FeCl3 can be reused as etching agents after adding 7–10 ml 30% solution of hydrogen peroxide per litre.  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