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 198 毫秒
1.
Novel poly(ether–imide)/polyurethane (PEI/PU)-based nanocomposite and foamed systems reinforced with graphene nanoplatelet (GNP) were developed. Field emission scanning electron microscopy revealed hexagonal nanocelluar morphology due to fine interaction between PEI/PU and functional GNP. Compression strength and modulus values were raised up to 72.3 MPa and 27.3 GPa, respectively, for PEI/PU/GNP Foam 1, thus revealing a defensive role of GNP layer against damage. Tmax of PEI/PU/GNP Foam 0.1–1 was measured as 479–565°C. The UL 94 showed V-0 rating for nanocomposite, while foams attained V-1 rating. Water absorption capacity was improved steadily with time and was at maximum after 96 h for PEI/PU/GNP Foam 1 (12.3%).  相似文献   

2.
High-performance gas separation membranes were fabricated using 0.5–3wt.% graphene nanoribbon (GNR). Poly(methyl methacrylate-co-methacrylic amide)-polyethylene glycol (PMMA-co-MA-PEG) copolymer was prepared via condensation and blended with polycarbonate to form PMMA-co-MA-PEG/PC. The PMMA-co-MA-PEG/PC and GNR-based nanocomposite possess seamless micro-branched morphological pattern. Tensile strength and Young’s Modulus of PMMA-co-MA-PEG/PC/GNR0.5-3 increased from 64.3–74.7MPa and 76.7–99.9MPa, respectively. GNR loading increased the permselectivity αCO2/N2 (25.4–41.6) of nanocomposite membrane relative to blend membrane (20.1). However, permeability PCO2 was decreased from 163.9 to 139.7 Barrer than blend (174.3 Barrer). PMMA-co-MA-PEG/PC/GNR revealed 51.6% increase and 24.7% decrease in permselectivity and permeability owing to molecular sieving and barricade characteristics of graphene nanoribbon.  相似文献   

3.
A concept of preparing high‐temperature proton exchange membranes with layer‐by‐layer (LBL) self‐assembly technique was proposed and the sulfonated polyetheretherketone (SPEEK) and polyurethane (PU) with 200 LBL deposition cycles denoting (SPEEK/PU)200 membrane was prepared in this research. Owing to the strong electrostatic interaction between ? group in SPEEK and ? C? N+ group in PU, (SPEEK/PU)200 membrane with LBL self‐assembly structure showed a favorable structural stability. The phosphoric acid (PA)‐doped (SPEEK/PU)200 membrane showed a higher proton conductivity relative to PA doped SPEEK/PU membrane by solution casting method (SPEEK/PU)200/40%PA membrane possessed a proton conductivity value of 2.90 × 10?2 S/cm at 150 °C under anhydrous conditions. The LBL self‐assembly structure provided a possibility to reduce the negative effect from polymer skeleton blocking charge carrier species even immobilizing protons. Moreover, the (SPEEK/PU)200 membrane presented the particularly noteworthy mechanical property even with PA doping. The tensile stress values at break were 72.8 and 24.1 MPa, respectively, for (SPEEK/PU)200 and (SPEEK/PU)200/40%PA membrane at room temperature, which were obviously higher than the reported values of 15.9 and 2.81 MPa for SPEEK/PU and SPEEK/PU/60%PA membrane. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 3446–3454  相似文献   

4.
A series of novel poly(urethane amide) films were prepared by the reaction of a polyurethane (PU) prepolymer and a soluble polyamide (PA) containing aliphatic hydroxyl groups in the backbone. The PU prepolymer was prepared by the reaction of polyester polyol and 2,4‐tolylenediisocyanate and then was end‐capped with phenol. Soluble PA was prepared by the reaction of 1‐(m‐aminophenyl)‐2‐(p‐aminophenyl)ethanol and terephthaloyl chloride. The PU prepolymer and PA were blended, and the clear, transparent solutions were cast on glass substrates; this was followed by thermal treatments at various temperatures to produce reactions between the isocyanate group of the PU prepolymer and the hydroxyl group of PA. The opaque poly(urethane amide) films showed various properties, from those of plastics to those of elastomers, depending on the ratio of the PU and PA components. Dynamic mechanical analysis showed two glass‐transition temperatures (Tg's), a lower Tg due to the PU component and a higher Tg due to the PA component, suggesting that the two polymer components were phase‐separated. The rubbery plateau region of the storage modulus for the elastic films was maintained up to about 250 °C, which is considerably higher than for conventional PUs. Tensile measurements of the elastic films of 90/10 PU/PA showed that the elongation was as high as 347%. This indicated that the alloying of PU with PA containing aliphatic hydroxyl groups in the backbone improved the high‐temperature properties of PU and, therefore, enhanced the use temperature of PU. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 3497–3503, 2002  相似文献   

5.
Two-dimensional (2D) transition-metal carbides, nitrides, and carbonitrides exhibit excellent thermal, mechanical, and electrical properties, and MXene displays good advantages in improving the mechanical properties of composites. In this study, the injection molding method is employed to introduce d-Ti3C2Tx MXene into the polyamide 66 matrix, yielding a d-Ti3C2Tx/PA66 nanocomposite. The tensile strength, flexural strength, Young's modulus, and hardness of d-Ti3C2Tx/PA66 nanocomposites are higher than that of pure PA66, owing to the high strength of the d-Ti3C2Tx nanosheets, good interfacial bonding, and stress transfer between the PA66 matrix and d-Ti3C2Tx nanofiller. TGA and DMA results revealed that the addition of d-Ti3C2Tx into the PA66 matrix improves the Tg, Es, creep resistance, and recovery properties, as well as the thermal stability of PA66 in an oxidizing atmosphere.  相似文献   

6.
The resin infiltration technique was used for impregnation of graphene oxide (GO) and graphene oxide-functional carbon nanotube (GO-ox-CNT) with polyvinyl chloride (PVC) and poly(styrene-co-maleic anhydride) (PSMA) blend. Two series of buckypapers were prepared with GO and GO-ox-CNT content. The morphology of PVC/PSMA/GO composite paper was porous, while PVC/PSMA/GO-ox-CNT revealed exclusive morphology with GO islands and intercalated ox-CNT networks covered with PVC/PSMA. Tmax of PVC/PSMA/GO-ox-CNT 0.1 was 571°C, whereas PVC/PSMA/GO 0.1 had a lower value (559°C). Tg of PVC/PSMA/GO-ox-CNT 0.1 was also higher (290°C). Cone calorimetric results showed a decrease in PHRR from 499 (blend) to 186 kW/m2 with 0.3 g GO-ox-CNT.  相似文献   

7.
The (p, ρ, T, x) properties for binary mixtures of CO2 (volume fraction purity 0.99999) and isobutane (mole fraction purity 0.99988) {x1 CO2 + x2 isobutane (x1 = 0.2482, 0.4718, and 0.7506)} were measured in the compressed liquid phase using a metal-bellows variable volumometer. Measurements were conducted from T = (280 to 440) K and (3 to 200) MPa. The expanded uncertainties (k = 2) were estimated to be: temperature, <3 mK; pressure, 1.5 kPa (p ? 7 MPa), 0.06% (7 MPa < p ? 50 MPa), 0.1% (50 MPa < p ? 150 MPa), 0.2% (p > 150 MPa); density, 0.10%; and composition, 4.4 · 10−4. At >100 MPa and T = (280 or 440) K, the uncertainties in the density measurements increased to 0.14% and 0.22%, respectively. The data are compared with the available equation of state. The excess molar volumes, , of the mixtures were calculated and plotted as a function of temperature and pressure.  相似文献   

8.
Silicone-modified graphene was successfully synthesized by treating graphene oxide with 3-aminopropyltriethoxysilane (AMEO) and then reduced by hydrazine hydrate. Subsequently, the AMEO-functionalized graphene was incorporated into polyurethane (PU) matrix to prepare AMEO-functionalized graphene/PU nanocomposite coatings. The functionalized graphene could disperse homogenously by means of a covalent connection with PU. AMEO-functionalized graphene (AFG)-reinforced PU nanocomposite coatings showed more excellent mechanical and thermal properties than those of pure PU. A 227 % increase in tensile strength and a 71.7 % improvement of elongation at break were obtained by addition 0.2 wt% of AFG. Meanwhile, thermogravimetric analysis reveals that thermal degradation temperature was enhanced almost 50 °C higher than that of neat PU, and differential scanning calorimetry analysis demonstrates that glass transition temperature decreased by around 9 °C. The thermal conductivity of AFG/PU nanocomposite coatings also increased by 40 % at low AFG loadings of 0.2 wt%.  相似文献   

9.
An in situ chemical synthesis approach has been developed to prepare SnO2–graphene nanocomposite. Field emission scanning electron microscopy and transmission electron microscopy observation revealed the homogeneous distribution of SnO2 nanoparticles (4–6 nm in size) on graphene matrix. The electrochemical reactivities of the SnO2–graphene nanocomposite as anode material were measured by cyclic voltammetry and galvanostatic charge/discharge cycling. The as-synthesized SnO2–graphene nanocomposite exhibited a reversible lithium storage capacity of 765 mAh/g in the first cycle and an enhanced cyclability, which can be ascribed to 3D architecture of the SnO2–graphene nanocomposite.  相似文献   

10.
Sulfur‐containing polyisobutylene (PIB)‐based polyurethane nanocomposite (PIBs‐PU/NC) was synthesized using HO? CH2CH2? S? PIB? S? CH2CH2? OH for the soft segment, conventional hard segments of MDI and BDO, and organically modified montmorillonite (OmMMT) nanolayers. The properties of PIBs‐PU/NC containing 72.5% PIB and 0.5% OmMMT were studied and contrasted with unmodified PIBs‐PU. PIBs‐PU/NC produces colorless optically clear films exhibiting enhanced tensile strength, elongation, oxidative–hydrolytic stability, and creep resistance relative to that of PIBs‐PU. FTIR spectroscopy indicates H bonded S atoms between soft and hard segments, and OmMMT nanolayers. DSC and XRD suggest randomly dispersed low‐periodicity crystals and urea groups between galleries. We propose that minute amounts of OmMMT nanolayers become covalently attached to polyurethane chains and beneficially affect properties by acting as co‐chain extender/reinforcing filler. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 2760–2765  相似文献   

11.
Abstract

Based on the aromatic diamine monomer containing di-tert-butylbenzene and methyl groups, this work proposes its polymerization with four different dicarboxylic acids. The prepared polyamides (PA 3a–3d) were characterized by GPC, FTIR, 1H NMR, mechanical, thermal, optical and gas separated techniques. They exhibited high solubility and good optical transparency. Their optical transmittance at 450?nm wavelength was in the range of 81.4%–86.8%, and the cutoff wavelength was in the range of 327–352?nm. The membranes also had good mechanical properties with tensile strength of 79.7–91.4?MPa, elongations at breaks of 9.0–10.9% and initial modulus of 1.5–1.9?GPa. Meanwhile, these membranes possessed good thermal properties with glass transition temperature (T g) values of 226–246?°C. The permeability of CH4, N2, and CO2 for these membranes was tested by constant pressure-variable volume method. The PA 3d containing tert-butyl moiety in the diacid units exhibited highest permeability (PCO2 = 31.4 and PN2 = 1.9) whereas PA 3c containing hexafluoroisopropylidene moiety exhibited highest selectivity (CO2/CH4 = 22.2).  相似文献   

12.
Poly(urethane‐benzoxazine) films as novel polyurethane ( PU )/phenolic resin composites were prepared by blending a benzoxazine monomer ( Ba ) and PU prepolymer that was synthesized from 2,4‐tolylene diisocyanate (TDI) and polyethylene adipate polyol (MW ca. 1000) in 2 : 1 molar ratio. DSC of PU/Ba blend showed an exotherm with maximum at ca. 246 °C due to the ring‐opening polymerization of Ba, giving phenolic OH functionalities that react with isocyanate groups in the PU prepolymer. The poly(urethane‐benzoxazine) films obtained by thermal cure were transparent, with color ranging from yellow to pale wine with increase of Ba content. All the films have only one glass transition temperature (Tg ) from viscoelastic measurements, indicating no phase separation in poly(urethane‐benzoxazine) due to in situ polymerization. The Tg increased with the increase of Ba content. The films containing 10 and 15% of Ba have characteristics of an elastomer, with elongation at break at 244 and 182%, respectively. These elastic films exhibit good resilience with excellent reinstating behavior. The films containing more than 20% of Ba have characteristics of plastics. The poly(urethane‐benzoxazine) films showed excellent resistance to the solvents such as tetrahydrofuran, N,N‐dimethyl formamide, and N‐methyl‐2‐pyrrolidinone that easily dissolve PU s. Thermal stability of PU was greatly enhanced even with the incorporation of a small amount of Ba . © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4165–4176, 2000  相似文献   

13.
This report describes a new route to covalently bonded polymer–graphene nanocomposites and the subsequent enhancement in thermal and mechanical properties of the resultant nanocomposites. At first, the graphite is oxidized by the modified Hummers method followed by functionalization with Octadecylamine (ODA). The ODA functionalized graphite oxides are reacted with methacryloyl chloride to incorporate polymerizable ? C?C? functionality at the nanographene platelet surfaces, which were subsequently employed in in situ polymerization of methylmethacrylate to obtain covalently bonded poly(methyl methacrylate) (PMMA)–graphene nanocomposites. The obtained nanocomposites show significant enhancement in thermal and mechanical properties compared with neat PMMA. Thus, even with 0.5 wt % graphene nanosheets, the Tg increased from 119 °C for neat PMMA to 131 °C for PMMA–graphene nanocomposite, and the respective storage modulus increased from 1.29 to 2 GPa. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 48: 4262–4267, 2010  相似文献   

14.
Nanocrystalline tin (Sn) compounds such as SnO2, SnS2, SnS, and graphene nanocomposites were prepared using hydrothermal method. The X-ray diffraction (XRD) pattern of the prepared nanocomposite reveals the presence of tetragonal SnO2, hexagonal SnS2, and orthorhombic SnS crystalline structure in the SnO2/graphene nanosheets (GNS), SnS2/GNS, and SnS/GNS nanocomposites, respectively. Raman spectroscopic studies of the nanocomposites confirm the existence of graphene in the nanocomposites. The transmission electron microscopy (TEM) images of the nanocomposites revealed the formation of homogeneous nanocrystalline SnO2, SnS2, and SnS particle. The weight ratio of graphene and Sn compound in the nanocomposite was estimated using thermogravimetric (TG) analysis. The cyclic voltammetry experiment shows the irreversible formation of Li2O and Li2S, and reversible lithium-ion (Li-ion) storage in Sn and GNS. The charge–discharge profile of the nanocomposite electrodes indicates the high capacity for the Li-ion storage, and the cycling study indicates the fast capacity fading due to the poor electrical conductivity of the nanocomposite electrodes. Hence, the ratio of Sn compounds (SnO2) and GNS have been altered. Among the examined SnO2:GNS nanocomposites ratios (35:65, 50:50, and 80:20), the nanocomposite 50:50wt% shows high Li-ion storage capacity (400 mAh/g after 25 cycles) and good cyclability. Thus, it is necessary to modify GNS and Sn compound composition in the nanocomposite to achieve good cyclability.  相似文献   

15.
Can commodity polymers are made to be healable just by blending with self-healable polymers? Here we report the first study on the fundamental aspect of this practically challenging issue. Poly(ether thiourea) (PTUEG3; Tg=27 °C) reported in 2018 is extraordinary in that it is mechanically robust but can self-heal even at 12 °C. In contrast, poly(octamethylene thiourea) (PTUC8; Tg=50 °C), an analogue of PTUEG3, cannot heal below 92 °C. We found that their polymer blend self-healed in a temperature range above 32 °C even when its PTUEG3 content was only 20 mol %. Unlike PTUEG3 alone, this polymer blend, upon exposure to high humidity, barely plasticized, keeping its excellent mechanical properties due to the non-hygroscopic nature of the PTUC8 component. CP/MAS 13C NMR analysis revealed that the polymer blend was nanophase-separated, which possibly accounts for why such a small amount of PTUEG3 provided the polymer blend with humidity-tolerant self-healable properties.  相似文献   

16.
Development of high‐strength hydrogels has recently attracted ever‐increasing attention. In this work, a new design strategy has been proposed to prepare graphene oxide (GO)/polyacrylamide (PAM)/aluminum ion (Al3+)‐cross‐linked carboxymethyl hemicellulose (Al‐CMH) nanocomposite hydrogels with very tough and elastic properties. GO/PAM/Al‐CMH hydrogels were synthesized by introducing graphene oxide (GO) into PAM/CMH hydrogel, followed by ionic cross‐linking of Al3+. The nanocomposite hydrogels were characterized by means of FTIR, X‐ray diffraction (XRD), and scanning electron microscopy/energy‐dispersive X‐ray analysis (SEM‐EDX) along with their swelling and mechanical properties. The maximum compressive strength and the Young's modulus of GO3.5/PAM/Al‐CMH0.45 hydrogel achieved values of up to 1.12 and 13.27 MPa, increased by approximately 6488 and 18330 % relative to the PAM hydrogel (0.017 and 0.072 MPa). The as‐prepared GO/PAM/Al‐CMH nanocomposite hydrogels possess high strength and great elasticity giving them potential in bioengineering and drug‐delivery system applications.  相似文献   

17.
The miscibility of polycarbonate (PC) with poly(methyl methacrylate-co-cyclohexyl methacrylate) (PMCHM) and with poly(methyl methacrylate) (PMMA) was studied by nuclear magnetic resonance (NMR) 1H spin-lattice relaxation time in the rotating frame (1H T1p), differential scanning calorimetry (DSC), and transmission electron microscopy (TEM). A blend of PC/PMCHM (50/50 wt/wt) with the acrylic component PMCHM, a copolymer of PMMA and poly(cyclohexyl methacrylate) (80/20 wt/wt), shows only one T1p value, which indicates high miscibility in this blend. A blend of PC/PMMA (50/50 wt/wt) shows two 1H T1p values, which are similar to those of the homopolymers PC and PMMA. These results indicate high immiscibility. The “domain size” calculated from NMR results of the miscible blend PC/PMCHM is approximately 40 Å. The results of DSC and TEM are similar to the NMR results. However, TEM results show the presence of 3% PC domains in the PC/PMCHM blend, which are not seen by NMR or DSC. Those PC domains are approximately 500 Å. A strong intramolecular repulsion in the copolymer PMCHM and specific intermolecular interactions between PC and PMMA may explain the miscibility in the PC/PMCHM system. © 1994 John Wiley & Sons, Inc.  相似文献   

18.
Nanostructured Fe2O3–graphene composite was successfully fabricated through a facile solution-based route under mild hydrothermal conditions. Well-crystalline Fe2O3 nanoparticles with 30–60?nm in size are highly encapsulated in graphene nanosheet matrix, as demonstrated by various characterization techniques. As electrode materials for supercapacitors, the as-obtained Fe2O3–graphene nanocomposite exhibits large specific capacitance (151.8?F?g?1 at 1?A?g?1), good rate capability (120?F?g?1 at 6?A?g?1), and excellent cyclability. The significantly enhanced electrochemical performance compared with pure graphene and Fe2O3 nanoparticles may be attributed to the positive synergetic effect between Fe2O3 and graphene. In virtue of their superior electrochemical performance, they will be promising electrode materials for high-performance supercapacitors applications.  相似文献   

19.
《先进技术聚合物》2018,29(1):528-540
Two series (random and block) poly(glycolide‐co‐ε‐caprolactone) macrodiols with various glycolide to ε‐caprolactone ratios (50/50 and 30/70, R‐PG50C, R‐PG30C, B‐PG50C, and B‐PG30C) were synthesized. Next, segmented polyurethanes (PUs) were synthesized based on the synthesized macrodiols, 1,6‐hexamethylene diisocyanate and 1,4‐butanediol (PU‐R30, PU‐R50, PU‐B30, and PU‐B50). Effect of glycolide (G) and ε‐caprolactone (C) monomers arrangement (random or block) on the PUs properties were investigated via FTIR, 1H NMR, DSC, TGA, DMA, SEM, and mechanical tests. All PUs illustrated Tg (−33°C to −48°C) and Tm (102°C to 139°C) corresponding to the soft and the hard segments, respectively. Polymers based on block macrodiols also showed Tm related to the soft segments. While PUs underwent a two‐step thermal degradation, the PUs based on block macrodiols indicated higher degradation temperature. Dynamic mechanical analysis results evidenced development of a well‐defined microphase separated structure in PU‐R30. Contact angle (about 70°‐80°) and water uptake (around 20% after 24 hours) of the PU films are close to those suitable for tissue engineering materials. The PU based on R‐PG30C (PU‐R30) exhibited the highest tensile strength (2.87 MPa) followed by PU‐B50 and PU‐R50. Over a 63‐day in vitro degradation study in phosphate buffered saline, the PUs showed variable weight loss (up to 40%) depending on their soft segments composition and arrangement. Also, the PUs showed no cytotoxicity. Thus, these PUs with tunable biodegradation rate and mechanical properties are suitable candidates for tissue engineering.  相似文献   

20.
This work studies the electrical, rheological, and thermal characteristics for polyurethane (PU) capped with tetraaniline as a new material, tetraaniline-containing poly(urethane–urea) (TAPU). The conductivities can be increased from less than 10−10 S/cm for pure PU to 10−4 S/cm for TAPU, independently of the length of the soft segment in the TAPU backbone chain. The tensile strength and modulus are increased when PU is copolymerized with tetraaniline. The viscoelastic creep can be effectively simulated using a Burgers model. Additionally, TAPU has higher viscosity, higher retardation time, and lower compliance J 1 than regular PU. Restated, TAPU exhibits less elastic but superior permanent deformation than PU because tetraaniline functions as a chain holder. The thermogravimetric analytic (TGA) results reveal that TAPU has lower T d, smaller T mw1 and T mw2, and higher char yield because the dehydration of the urea-containing polymer produces a thin layer from a nitrogen compound on the polymer’s surface, which insulates the underlying polymer from heat and oxygen.  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