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 125 毫秒
1.
Jing Quan  Qi Wu  Xian-Fu Lin 《Polymer》2008,49(16):3444-3449
Optically active polymeric prodrugs containing saccharides were prepared via a facile procedure combining enzymatic resolution with chemical polymerization. Polymerizable optically active chlorphenesin derivatives were obtained in excellent optically purity (ee > 99.9%) and high yield (∼50%) in short time (∼4 h) by lipase-catalyzed resolution after optimization of reaction conditions. Then, polymerizable optically active monomer was copolymerized with different polymerizable glycolipids using free radical polymerization method. The obtained optically active polymeric prodrugs bearing (R)-chlorphenesin residue were characterized by IR, NMR and GPC. In vitro released studies showed that the cumulative released optically pure chlorphenesin (ee = 88-92%) from the polymeric prodrug (VADG) was from 19.3% to 34.3% in pH 7.4, pH 5.4 and pH 1.2 after 7 days. The cumulative liberation rate of optically active polymeric prodrugs (ee = ∼88%) with different glycolipids was from 26.4% to 38.5% in pH 1.2.  相似文献   

2.
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   

3.
A star polymer was synthesized by addition of 1,4-diethynyl-2,5-dimethylbenzene as linking agent (30 °C, 24 h) after living polymerization of [(o-trifluoromethyl)phenyl]acetylene (o-CF3PA) with MoOCl4-n-Bu4Sn-EtOH catalyst (in anisole, 30 °C, 20 min; [Mo]=10 mM, [P]/[Mo]=40%, [o-CF3PA]0=200 mM). The Mn values of the living and star polymers were 8.1×103 and 5.3×104, respectively, according to gel permeation chromatography, while these values determined by multi-angle laser light scattering (MALLS) were 7.8×103 and 2.5×105. The Mw/Mn and arm number of the star polymer were 1.04 and 29, respectively, according to MALLS. The molecular weight and arm number of star polymer increased with increasing linking agent concentration and polymerization temperature.  相似文献   

4.
Thermal polymerization of methyl (meth)acrylate (MMA) was carried out using 2-cyanoprop-2-yl-1-dithionaphthalate (CPDN) and cumyl dithionaphthalenoate (CDN) as chain transfer agents. The kinetic study showed the existence of induction period and rate retardation, especially in the CDN mediated systems. The molecular weights of the polymers increased linearly with the monomer conversion, and the molecular weight distributions (Mw/Mns) of the polymers were relatively narrow up to high conversions. The maximum number-average molecular weights (Mns) reached to 351?900 g/mol (Mw/Mn = 1.47) and 442?400 g/mol (Mw/Mn = 1.29) in the systems mediated by CPDN and CDN, respectively. Chain-extension reactions were also successfully carried out to obtain higher molecular weight PMMA and PMMA-block-polystyrene (PMMA-b-PSt) copolymer with controlled structure and narrow Mw/Mn. Thermal polymerization of methyl acrylate (MA) in the presence of CPDN, or benzyl (2-phenyl)-1-imidazolecarbodithioate (BPIC) also demonstrated “living”/controlled features with the experimented maximum molecular weight 312?500 g/mol (Mw/Mn = 1.57). The possible initiation mechanism of the thermal polymerization was discussed.  相似文献   

5.
Liang Tong 《Polymer》2008,49(21):4534-4540
Perfluorocyclobutyl aryl ether-based amphiphilic diblock copolymer containing hydrophilic poly(ethylene glycol) segment was synthesized by atom transfer radical polymerization (ATRP). Perfluorocyclobutyl-containing methacrylate-based monomer, 4-(4′-p-tolyloxyperfluorocyclobutoxy)benzyl methacrylate, was prepared firstly, which can be polymerized by ATRP in a controlled way to obtain well-defined homopolymers with narrow molecular weight distributions (Mw/Mn ≤ 1.30). The molecular weights increased linearly with the conversions of monomer and the apparent polymerization rate exhibited first-order relation with respect to the concentration of monomer. ATRP of 4-(4′-p-tolyloxyperfluorocyclobutoxy)benzyl methacrylate was initiated by PEG-based macroinitiators with different molecular weights to obtain amphiphilic diblock copolymers with narrow molecular weight distributions (Mw/Mn < 1.35) and the number of perfluorocyclobutyl linkage can be tuned by the feed ratio and the conversion of the fluorine-containing methacrylate monomer. The critical micelle concentrations of these amphiphilic diblock copolymers in water and brine were determined by fluorescence probe technique. The morphologies of the micelles were found to be spheres by TEM.  相似文献   

6.
Akito Fukui 《Polymer》2009,50(17):4159-5967
Diarylacetylenes having fluorenyl groups and other substituents (trimethylsilyl, t-butyl, bromine, fluorine) (1a-1) were polymerized with TaCl5-n-Bu4Sn. Monomers 1a-l produced high molecular weight polymers 2a-l (Mw 5.1 × 105-1.3 × 106) in 12-59% yields. All of the polymers were soluble in common organic solvents, and gave tough free-standing membranes by the solution casting method. The onset temperatures of weight loss of polymers 2a-l in air were over 400 °C, indicating considerably high thermal stability. All the polymer membranes showed high gas permeability; e.g., the oxygen permeability coefficient (PO2) of 2a was as large as 4800 barrers. Membrane 2d possessing two fluorine atoms at meta and para positions of the phenyl ring showed the highest oxygen permeability (PO2 = 6600 barrers) among the present polymers.  相似文献   

7.
The controlled polymerization of vinyl chloride (VC) with tert-butyllithium (tert-BuLi) was investigated. The polymerization of VC with tert-BuLi at −30 °C proceeded to give a high molecular weight polymer in good yield. In the polymerization of VC −30 to 0 °C under nearly bulk, the relationship between the Mn of polymers and polymer yields gave a straight line passed through the origin, but the Mw/Mn of PVC was not narrow. When CH2Cl2 was used as polymerization solvent, the Mn of PVC increased with the polymer yield, and the Mw/Mn of 1.25 was obtained. Structure analysis of the resulting polymers indicates that the main chain structure could be regulated in the polymerization of VC with tert-BuLi. Accordingly, a control of molecular weight of polymer and main chain structure is possible in the polymerization of VC with tert-BuLi.  相似文献   

8.
Phenanthrene α-end-labeled poly(N-decylacrylamide-b-N,N-diethylacrylamide) (PDcAn-b-PDEAm) block copolymers consisting in a highly hydrophobic block (n = 11) and a thermoresponsive block with variable length (79 ≤ m ≤ 468) were synthesized by reversible addition-fragmentation chain transfer (RAFT) polymerization. A new phenanthrene-labeled chain transfer agent (CTA) was synthesized and used to control the RAFT polymerization of a hydrophobic acrylamide derivative, N-decylacrylamide (DcA). This first block was further used as macroCTA to polymerize N,N-diethylacrylamide (DEA) in order to prepare diblock copolymers with the same hydrophobic block of PDcA (number average molecular weight: Mn = 2720 g mol−1, polydispersity index: Mw/Mn = 1.13) and various PDEA blocks of several lengths (Mn = 10,000-60,000 g mol−1) with a very high blocking efficiency. The resulting copolymers self-assemble in water forming thermoresponsive micelles. The critical micelle concentration (CMC) was determined using Förster resonance energy transfer (FRET) between phenanthrene linked at the end of the PDcA block and anthracene added to the solution at a low concentration (10−5 M), based on the fact that energy transfer only occurs when phenanthrene and anthracene are located in the core of the micelle. The CMC (∼2 μM) was obtained at the polymer concentration where the anthracene fluorescence intensity starts to increase. The size of the polymer micelles decreases with temperature increase around the lower critical solution temperature of PDEA in water (LCST ∼ 32 °C) owing to the thermoresponsiveness of the PDEA shell.  相似文献   

9.
Vinyl ester (VE) monomers with bimodal molecular weight distributions were prepared by reacting methacrylic acid with blends of monodisperse epoxy resins ranging in molecular weight from 350-7000 g/mol. Monodisperse vinyl ester monomers were prepared from epoxy resins of a single molecular weight. The extent of vinyl ester formation was found to be near complete and side reactions, such as etherification, did not occur to a significant extent. The viscosities of these vinyl ester resins were measured as a function of styrene content. It was found that resin viscosity, η, increased exponentially and predictably as both the styrene content (S) decreased and as the number average molecular weight (Mn) of the vinyl ester monomers increased: η∼exp(Mn)/exp(S). Cure kinetics studies showed that the vinyl ester reactivity ratio decreased to 0.1 from 0.6 for bimodal blends relative to monodisperse resins while the styrene reactivity ratio increased from 0.4 to 0.6. Thus, the microgels in bimodal blends were smaller than in monodisperse resins. Emissions studies proved that decreasing the styrene content reduced the VOC emission rate and total emissions. Higher VE molecular weights decreased the overall emissions due to a reduction in monomer mobility. Tg decreased from 143 to 125 °C as Mn of the VE monomers increased from 540 to 920 g/mol; yet, Tg of these bimodal blends were still equal to or greater than that of commercial VE resins (∼125 °C). The fracture toughness of bimodal blends increased from ∼100 to ∼330 J/m2 as VE Mn increased from 540 to 920 g/mol because of matrix toughening. The fracture properties did not improve as the styrene content increased from 35 to 45 wt% because of corresponding changes in the morphology. Yet, there were numerous low VOC bimodal formulations with fracture properties in excess of the low VOC Dow Derakane 441-400 (110 J/m2) and even the industry standard Derakane 411-350 (240 J/m2).  相似文献   

10.
Isotactic polypropylene (iPP)-polystyrene (PS) and iPP-poly(methyl methacrylate) (PMMA) multiblock copolymers were synthesized by atom transfer radical coupling (ATRC) of PS-iPP-PS and PMMA-iPP-PMMA triblock copolymers obtained by atom transfer radical polymerization (ATRP) of styrene (St) and methyl methacrylate (MMA), respectively, using α,ω-dibromoisobutyrateoligopropylene (iPP-Br) as a bifunctional macroinitiator. The iPP-Br was prepared by hydroxylation and subsequent esterification of telechelic oligopropylene having terminal vinylidene double bonds at both ends obtained by controlled thermal degradation of iPP. ATRP of St and (meth) acrylic monomers using iPP-Br formed the corresponding triblock copolymers. It was confirmed that the PMMA-iPP-PMMA triblock copolymer was effective as the compatibilizer for the iPP/PMMA blend. An iPP-PS multiblock copolymer (Mn: 25?000 g/mol and Mw/Mn: 4.1) was prepared by ATRC of PS-iPP-PS triblock copolymer (Mn: 8900 g/mol and Mw/Mn: 1.3). ATRC with St of PMMA-iPP-PMMA triblock copolymer (Mn: 13?000 g/mol and Mw/Mn: 1.4) provided an iPP-PMMA multiblock copolymer containing St chains (Mn: 39?000 g/mol and Mw/Mn: 2.8).  相似文献   

11.
A series of poly(ω-pentadecalactone) (PPDL) samples, synthesized by lipase catalysis, were prepared by systematic variation of reaction time and water content. These samples possessed weight-average molecular weights (Mw), determined by multi-angle laser light scattering (MALLS), from 2.5 × 104 to 48.1 × 104. Cold-drawing tensile tests at room temperature of PPDL samples with Mw between 4.5 × 104 and 8.1 × 104 showed a brittle-to-ductile transition. For PPDL with Mw of 8.1 × 104, inter-fibrillar slippage dominates during deformation until fracture. Increasing Mw above 18.9 × 104 resulted in enhanced entanglement network strength and strain-hardening. The high Mw samples also exhibited tough properties with elongation at break about 650% and tensile strength about 60.8 MPa, comparable to linear high density polyethylene (HDPE). Relationships among molecular weight, Young's modulus, stress, strain at yield, melting and crystallization enthalpy (by differential scanning calorimetry, DSC) and crystallinity (from wide-angle X-ray diffraction, WAXD) were correlated for PPDL samples. Similarities and differences of linear HDPE and PPDL molecular weight dependence on their mechanical and thermal properties were also compared.  相似文献   

12.
Fabio Fabri  Wanda de Oliveira 《Polymer》2006,47(13):4544-4548
Half-sandwich samarium(III) diketiminate bromide was successfully synthesized and was shown to be active in methyl methacrylate (MMA) polymerization. The effects of temperature, polymerization time and catalyst concentration were studied. Activities of ca. 18 kg of polymethacrylate (PMMA) per mol of samarium per hour were obtained under optimum conditions (0 °C and a MMA/catalyst molar ratio of 100/1), giving a polymer with a molar mass Mn>24,000 g mol−1 and a molar mass distribution (Mw/Mn)<1.4. After 1 h of polymerization, conversions of MMA as high as 96% were observed.  相似文献   

13.
Lihui Cao  Weimin Dong  Xuequan Zhang 《Polymer》2007,48(9):2475-2480
The oxovanadium phosphonates (VO(P204)2 and VO(P507)2) activated by various alkylaluminums (AlR3, R = Et, i-Bu, n-Oct; HAlR2, R = Et, i-Bu) were examined in butadiene (Bd) polymerization. Both VO(P204)2 and VO(P507)2 showed higher activity than those of classical vanadium-based catalysts (e.g. VOCl3, V(acac)3). Among the examined catalysts, the VO(P204)2/Al(Oct)3 system (I) revealed the highest catalytic activity, giving the poly(Bd) bearing Mn of 3.76 × 104 g/mol, and Mw/Mn ratio of 2.9, when the [Al]/[V] molar ratio was 4.0 at 40 °C. The polymerization rate for I is of the first order with respect to the concentration of monomer. High thermal stability of I was found, since a fairly good catalytic activity was achieved even at 70 °C (polymer yield > 33%); the Mn value and Mw/Mn ratio were independent of polymerization temperature in the range of 40-70 °C. By IR and DSC, the poly(Bd)s obtained had high 1,2-unit content (>65%) with atactic configuration. The 1,2-unit content of the polymers obtained by I was nearly unchanged, regardless of variation of reaction conditions, i.e. [Al]/[V], ageing time, and reaction temperature, indicating the high stability of stereospecificity of the active sites.  相似文献   

14.
The conformation of single poly(methyl methacrylate) chain in the uniaxially stretched film was observed by the combination of the fluorescence labeling technique and scanning near-field optical microscopy (SNOM). The dimension of the individual probe chain (Mw = 1.99 × 106) along the extension axis was evaluated from SNOM image. In the high molecular weight matrix (Mw = 1.89 × 106) the average extension ratio at the single chain level coincided with the macroscopic extension ratio. The distribution of the chain conformation was in good agreement with that of the freely jointed chain followed by affine deformation. On the other hand, the probe chain embedded in the low molecular weight matrix (Mw = 1.76 × 105) showed the smaller extension than that expected from the affine deformation. This suggests that the conformation of the probe chain is affected by the relaxation of the short surrounding chains through disentanglement.  相似文献   

15.
Titanium complexes having tridentate triamine of the type N[CH2CH(Ph)(Ts)N]22− in combination with methylaluminoxane (MAO) was able to polymerize ethyl vinyl ether in good yields. The polymers obtained in general were having molecular weight in the order of 105 with narrow molecular weight distributions. Polymerization conditions had an impact on the molecular weight and the polydispersity index (PDI). Using chlorobenzene as the solvent the polymer had an Mn of 350?000 and PDI of 1.21, where as under neat conditions the Mn was 255?000 with PDI of 1.21. The type of solvent and the temperature dictated the polymerization rate and the polymer stereo regularity. The molecular weight of the polymer is distinctly governed by the polymerization temperature. Temperature ranging between −50 and ambient (30 °C) resulted in high molecular weight polymers and vice versa at a temperature of 60-70 °C resulted in low molecular weight polymers in moderate yields. The polymers obtained below 30 °C are highly stereo-regular compared to that of the ones produced at and above ambient temperature. The polymerization of iso-butyl vinyl ether (IBVE) was faster than that of linearly substituted n-butyl vinyl ether (BVE) and less bulky ethyl vinyl ether (EVE). The order of isotacticities of the polymers obtained are polyIBVE > polyBVE > polyEVE. The use of borate cocatalyst for activation generated narrow molecular weight polymers with a linear increase in the yield and molecular weight over time suggesting the living nature of the catalyst system.  相似文献   

16.
Weipu Zhu 《Polymer》2005,46(19):8379-8385
Rare earth (Nd, Y, La) p-tert-butylcalix[n]arene (n=4, 6, and 8) complexes alone have been developed to catalyze random and block copolymerizations of trimethylene carbonate (TMC) and 2,2-dimethyltrimethylene carbonate (DTC). The random or block structure and thermal behavior of the copolymers have been characterized by SEC, NMR, DSC, XRD and PLM. Random copolymer with Mw of 14,100 and Mw/Mn of 1.36 was prepared by neodymium p-tert-butylcalix[6]arene complex under the conditions: [TMC+DTC]0/[Nd]=400, 80 °C, 8 h. The reactivity ratios of TMC and DTC are measured to be rTMC=4.68 and rDTC=0.163, respectively. Random copolymerization could be well designed by the feeding ratio to prepare copolymers with controlled Tm and Tg. Only 8% TMC units in the polymer chain destroyed the crystallization of PDTC showing no Tm of the copolymer in the DSC analysis.  相似文献   

17.
Durairaj Baskaran 《Polymer》2003,44(8):2213-2220
Hyperbranched polymers were synthesized using anionic self-condensing vinyl polymerization (ASCVP) by forming ‘inimer’ (initiator within a monomer) in situ from divinylbenzene (DVB) and 1,3-diisopropenylbenzene (DIPB) using anionic initiators in THF at −40 °C. The reaction of equimolar amounts of DVB and nBuLi results in the formation of hyperbranched poly(divinylbenzene) through self-condensing vinyl polymerization (SCVP). The hyperbranched polymers were invariably contaminated with small amount of gel (<15%). No gelation was observed when using DIBP with anionic initiators. The presence of monomer-polymer equilibrium in the SCVP of DIPB restricts the growth of hyperbranched poly(DIPB). The inimer synthesized from DIPB at 35 °C undergoes intermolecular self-condensation to different extent depending on the nature of anionic initiator at −40 °C. The molecular weight of the hyperbranched polymers was higher when DPHLi was used as initiator. A small amount of styrene ([styrene]/[Li+]=1) was used to promote the chain growth by inducing cross-over reaction with styrene, and subsequent reaction of styryl anion with isopropenyl groups of inimer/hyperbranched oligomer. The hyperbranched polymers were soluble in organic solvents and exhibited broad molecular weight distribution (2<Mw/Mn<17).  相似文献   

18.
Wallace W. Yau 《Polymer》2007,48(8):2362-2370
Model calculations were performed to investigate the sensitivity of zero-shear melt viscosity (η0 or Eta0) on the molecular weight (MW) polydispersity of linear polymers. Simulated MW distributions (MWD) were generated with the generalized exponential (GEX) distribution function for various levels of polydispersity Mw/Mn and Mz/Mw. For linear entangled polymeric chains in the melt, the linear viscoelastic properties were predicted by using the double reptation blending rule and the so-called BSW relaxation time spectrum, named after the authors: Baumgaertel, Schausberger and Winter [Baumgaertel M, Schausberger A, Winter HH. Rheol Acta 1990;29:400-8]. Published rheological parameters appropriate for polyethylene were used in the calculations. It was found that Eta0 depended mostly on Mw, but it also significantly depended on the extent of high-MW polydispersity Mz/Mw. A revision to the fundamental MW dependency of Eta0 was proposed to compensate for this polydispersity effect. To offset the polymer polydispersity differences, we propose a new MW average (MHV or Mx with x = 1.5) to replace Mw in the historical rheological power-law equation of Eta0 ∝ Mwa, where the literature value of exponent “a” ranges from 3.2 to 3.6. The use of MHV instead of Mw in the power-law equation made the calculated Eta0 independent of the sample high-MW polydispersity. With the removal of the complication from polydispersity effect, the new Eta0 power law can now provide a more robust base for studying polymer long-chain branching (LCB). A new LCB index is thus proposed based on this new melt-viscosity power law. The values of MHV in the new power law can be calculated for polymer samples from the conventional gel permeation chromatographic (GPC) slice data.  相似文献   

19.
Zhicheng Xiao 《Polymer》2007,48(18):5388-5397
Small angle light scattering has been used to probe structure formation during isothermal crystallization of an ethylene-1-hexene copolymer (EH064, Mw = 70,000 g/mol, ρ = 0.900 g/cm3, Mw/Mn ∼ 2, 6.4 mol% hexene). It is shown that clear structural information on size scales ranging from hundreds of nanometers to several micrometers during early stage crystallization can be obtained by this method when crystallizing the polyethylenes at the high temperatures (above the peak melting temperature of a rapidly crystallized polymer sample) required for resolving early stage crystallization without the influence of the crystal growth. The results show that the early stage crystallization is characterized by large scale orientation fluctuations that precede the formation of local crystalline order manifest in X-ray scattering and the initial collapse of these large scale anisotropic/ordered domains. The scattering intensity increases exponentially with time initially, and the wave vector dependence of the growth rate of fluctuations is consistent with predictions for initial stages of a phase transformation process. However, the detailed mechanism cannot be described by existing models. The implications of our results are discussed within the context of proposed models for early stage crystallization.  相似文献   

20.
Ren-Shen Lee  Hua-Rong Li  Fu-Yuan Tsai 《Polymer》2005,46(24):10718-10726
A series of novel types of diblock poly(trans-4-hydroxy-N-benzyloxycarbonyl-l-proline)-block-poly(ε-caprolactone) (PHpr10-b-PCL) copolymers were synthesized by ring-opening polymerization from macroinitiator poly(trans-4-hydroxy-N-benzyloxycarbonyl-l-proline) (PHpr10) and ε-caprolactone (ε-CL) in the presence of organocatalyst dl-lactic acid (dl-LA). The Mn of the copolymers increased from 3370 to 19,040 g mol−1 with the molar ratio (10-100) of ε-CL to PHpr10. These products were characterized by differential scanning calorimetry (DSC), 1H NMR, and gel permeation chromatography. According to DSC, the glass-transition temperature (Tg) of the diblock copolymers depend on the molar ratio of monomer/initiator that were added. The hydrolytic degradation behavior of PHpr-b-PCLs was evaluated from weight-loss measurements and the change of Mn and Mw/Mn. With higher PCL contents resulted in a slower weight loss, while having a higher molecular weight loss percentage. Their micellar characteristics in an aqueous phase were investigated by fluorescence spectroscopy, transmission electron microscopy (TEM), and dynamic light scattering (DLS). The block copolymers formed micelles in the aqueous phase with critical micelle concentrations (CMCs) in the range of 1.33-4.22 mg L−1. The micelles exhibited a spindly shape and showed a narrow monodisperse size distribution. The obtained micelles have a relatively high drug-loading of about 26% when the feed weight ratio of amitriptyline hydrochloride (AM) to polymer was 1/1. An increase of molecular weight and hydrophobic components in copolymers produced a higher CMC value and greater loading efficiencies were observed.  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