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 484 毫秒
1.
Abstract

Films of high‐molecular‐weight amorphous polystyrene (PS, M w = 225 kg/mol, M w/M n = 3, T g‐bulk = 97°C, where T g‐bulk is the glass transition temperature of the bulk sample) and poly(methyl methacrylate) (PMMA, M w = 87 kg/mol, M w/M n = 2, T g‐bulk = 109°C) were brought into contact in a lap‐shear joint geometry at a constant healing temperature T h, between 44°C and 114°C, for 1 or 24 hr and submitted to tensile loading on an Instron tester at ambient temperature. The development of the lap‐shear strength σ at an incompatible PS–PMMA interface has been followed in regard to those at compatible PS–PS and PMMA–PMMA interfaces. The values of strength for the incompatible PS–PMMA and compatible PMMA–PMMA interfaces were found to be close, both being smaller by a factor of 2 to 3 than the values of σ for the PS–PS interface developed after healing at the same conditions. This observation suggests that the development of the interfacial structure at the PS–PMMA interface is controlled by the slow component, i.e., PMMA. Bonding at the three interfaces investigated was mechanically detected after healing for 24 hr at T h = 44°C, i.e., well below T g‐bulks of PS and PMMA, with the observation of very close values of the lap‐shear strength for the three interfaces considered, 0.11–0.13 MPa. This result indicates that the incompatibility between the chain segments of PS and PMMA plays a negligible negative role in the interfacial bonding well below T g‐bulk.  相似文献   

2.
This effort reports on novel fluorinated polyamide (FPA) and polyamide 1010 (PA1010)-based blends and graphene reinforced nanocomposite. PA1010/FPA (80:20) blend was opted as matrix material on the basis of molecular weight, thermal, and shear stress performance. Graphene was obtained through in situ chemical method of graphene oxide reduction. PA1010/FPA/Graphene nanocomposites was developed using various graphene loadings (up to 5 wt.%). Thin film coatings were prepared on glass substrate. Consequently, the PA1010/FPA/Graphene attained regular spongy morphological pattern. PA1010/FPA/Graphene 3 also showed improved T0 and Tmax of 534 and 591 °C relative to the neat blend (T10 423 °C; Tmax 551 °C). Limiting oxygen index measurement indicated better non-flammability of PA1010/FPA/Graphene 1–3 nanocomposite series (57–60%) relative to the blend series (28–31%). UL94 tests also showed V-0 rating for nanocomposites. Furthermore, PA1010/FPA/Graphene 3 nanocomposite revealed significantly high tensile strength (62 MPa), flexural modulus (1690 MPa), and adhesive properties to be utilized as coating materials. The nanocomposite coatings also displayed outstanding barrier properties against O2 and H2O compared with neat blends.  相似文献   

3.
Increase of photoluminescence (PL) from fullerenes (C60 and C70)-doped poly(alkyl methacrylate), such as poly(ethyl methacrylate) (PEMA), poly(isopropyl methacrylate) (PiPMA) and poly(isobutyl methacrylate) (PiBMA), have been studied under laser irradiation with wavelength of in air. After laser irradiation, PL peaks of all fullerenes doped-polymers are broadened and blue-shifted. This PL increase depends on the fullerene concentrations. By comparing with fullerenes-doped PMMA, fullerenes-doped PEMA have the greatest PL increase among the four kinds of polymers, including PEMA, PiPMA, PiBMA and PMMA. PL intensity of C70-doped polymers increases much more quickly than the corresponded C60-doped polymers at the initial stage of laser irradiation. Great change on their UV-visible absorption spectra before and after laser irradiation indicate some great variation on chemical structure of fullerene molecules dispersed in polymer matrix under laser irradiation. This great PL increase may be attributed to formation of fullerene oxide-polymer and oxidized fullerene-polymer adducts due to laser-induced photochemical reactions among fullerene, oxygen and polymer.  相似文献   

4.
The amorphous polymer surfaces of polystyrene (PS, M n=200 kg/mol, M w/M n=1.05) and poly(methyl methacrylate) (PMMA, M n=51.9 kg/mol, M w/M n≤1.07) were brought into contact at 21°C to form PS‐PS (for 54 days) and PMMA‐PMMA auto‐adhesive joints (for 11 days). After contact at that temperature corresponding to T g‐bulk ?81°C for PS and to T g‐bulk–88°C for PMMA, where T g‐bulk is the calorimetric glass transition temperature of the bulk sample, the bonded interfaces were fractured and their surfaces were analyzed by atomic force microscopy (AFM). The surface roughness, R q, of the fractured interfaces was larger by a factor of 3–4 than was that of the free PS and PMMA surfaces aged for the same period of time. A similar increase in R q was found by comparison of the free PS surface aged at T g‐bulk+15°C for 1 h and of the surface of the PS‐PS interface fractured after healing at T g‐bulk+15°C for 1 h. These observations, indicative of the deformation of the fractured interfaces, suggest the occurrence of some mass transfer across the interface even below T g‐bulk ?80°C.  相似文献   

5.
A novel organic/inorganic composite based on LiNi–ferrospinel with poly(methyl methacrylate) (PMMA) and polyaniline (PANI), PANI/PMMA/LiNi0.5Fe2O4 composite, was synthesized via a facile in-situ polymerization process. The structures of the resulting samples were investigated by X-ray diffraction, Fourier transform infrared spectroscopy, and atomic force microscopy. The optical and thermal properties of the PANI/PMMA/LiNi0.5Fe2O4 composite were studied by fluorescent spectroscopy and thermogravimetry analysis. It was indicated that the existence of LiNi0.5Fe2O4 (LFNO) in the PANI/PMMA/LFNO composite resulted in changes in the fluorescence spectra. The as-obtained composite may have potential for electrical and electromagnetic applications in antistatic materials, electromagnetic shields, radar absorbers, and so forth.  相似文献   

6.
Tao Sun  Jiayu Yu  Qi Yang  Jinxin Ma 《Ionics》2017,23(5):1059-1066
Cu-supported SnO2@C composite coatings constructed by interconnected carbon-based porous branches were fabricated by annealing Cu foils with films formed by knife coating DMF solution containing SnCl2, polyacrylonitrile (PAN), and poly(methyl methacrylate) (PMMA) on their surface in vacuum. The carbon-based porous branches consist of amorphous carbon matrices, SnO2 nanoparticles with a size of 30–100 nm mainly encapsulated inside, and many micropores with a size of 1–5 nm. The three-dimensional (3D) porous network structures of the SnO2@C composite were achieved by volatilization of PMMA and pyrolysis of SnCl2. The SnO2@C composite coatings demonstrate good cyclic performance with a high reversible capacity of 642 mA h g?1 after 100 cycles at a current density of 50 mA g?1 without apparent capacity fading during cycling and excellent rate performance with a capacity of 276 mA h g?1 at a high current density up to 10 A g?1.  相似文献   

7.
Nanocomposite polyurethane foams filled with different loadings (0.1–0.7 wt.%) of nanosized silica (average grain size of about 7 or 12 nm) and organoclay were prepared by a prepolymer method, and their mechanical properties were investigated. Statistical analysis of the size distribution of the foam cells was successfully applied for the characterization of their morphology. It was shown that the developed approach provided detailed analysis of the morphology development in PU foams, including the primary cell formation and their break-up and coalescence. The degree of phase separation in nanocomposite polyurethane foams in its dependence on nanofiller type and content was calculated from the IR spectra. The presence of silica nanoparticles and organoclays gives rise to significant differences in the mechanical (stress–strain) properties of the nanocomposite polyurethane foams with respect to the pure polymer.  相似文献   

8.
Novel composite materials are synthesized by incorporating N-acryloylmorpholine(ACMO) in highly concentrated phenanthrenequinone(PQ) doped poly(methyl methacrylate)(PMMA). The photosensitizer concentration of PQ was increased from 0.7 wt. % to 1.8 wt. %. The doping of ACMO component results in a higher diffraction efficiency and photosensitivity than a typical PQ/PMMA system. The enhanced performance of the material may stem from the ACMO molecules, which might open a new route for improving the holographic performance of the PQ/PMMA photopolymer.  相似文献   

9.
Effects of graphene nanoplatelet (GNP) addition on the electrical conductivity and optical absorbance of poly(methyl methacrylate)/graphene nanoplatelet (PMMA/GNP) composite films were studied. Optical absorbance and two point probe resistivity techniques were used to determine the variations of the optical and electrical properties of the composites, respectively. Absorbance intensity, A, and surface resistivity, Rs, of the composite films were monitored as a function of GNP mass fraction (M) at room temperature. Absorbance intensity values of the composites were increased and surface resistivity values were decreased by increasing the content of GNP in the composite. Electrical and optical percolation thresholds of composite films were determined as Mσ = 27.5 wt.% and Mop = 26.6 wt.%, respectively. The conductivity and the optical results were attributed to the classical and site percolation theories, respectively. Optical (βop) and electrical (βσ) critical exponents were calculated as 0.40 and 1.71, respectively.  相似文献   

10.
The molecular dynamics simulation (MD) was carried out to investigate the mechanical properties of pristine polymethylmethacrylate (PMMA) and the composites of PMMA mixed with the silver nanoparticles (PMMA/AgNPs) at two AgNP weight fractions at 0.60 and 1.77 wt%. From the stress–strain profiles by the tensile process, it can be seen that the improvement on Young’s modulus is insignificant at these lower AgNP fractions. The tensile strength of pristine PMMA can be slightly improved by the embedded AgNPs at 1.77 wt%, because the local density and strength of PMMA in the vicinity of AgNP surface within about 8.2 Å are improved. For the temperature effect on the mechanical properties of pristine PMMA and PMMA/AgNP composite, the Young’s moduli and strength of pristine PMMA and PMMA/AgNP composite significantly decrease at temperatures of 450 and 550 K, which are close to the predicted melting temperature of pristine PMMA about 460 K. At these temperatures, the PMMA materials become more ductile and the AgNPs within the PMMA matrix display higher mobility than those at 300 K. When the tensile strain increases, the AgNPs tend to get closer and the fracture appears at the PMMA part, leading to the close values of Young’s modulus and ultimate strength for pristine PMMA and PMMA/AgNP composite at 450 and 550 K.
Graphical abstract Stress–strain curves of pristine PMMA, polymethylmethacrylate (PMMA)/silver nanoparticles (AgNP) (0.60%), and PMMA/AgNP (1.77%). Inset images: local shear strain of pristine PMMA (red) and PMMA/AgNP (1.77%) (green).
  相似文献   

11.
The modification of graphite used in diamond synthesis with low concentrations of the fullerene C60-C70 extract (from 0.045 to 0.225 wt % of graphite mass) in the presence of a Ni-Mn metal catalyst at a pressure of 5 GPa in the temperature range 1600–1800 K is found to decrease the activation energy of the graphite-diamond phase transition from 160 ± 40 to 100 ± 40 kJ/mol.  相似文献   

12.
The present article has reported the effects of several nanofiller’s aspect ratio, length and interfacial strength on Mode-I fracture toughness (KIC) of geopolymer as the matrix of continuous fibre reinforced composites. These nanofillers have been chosen based on the variations in the surface chemistry and nature of interfacial bonding with geopolymer, which include Carbon, Alumina and Silicon carbide. Geopolymer matrix was subjected to the addition of single volume fraction, 2% of each type of nanofiller with two aspect ratios, designated as nanoparticles and nanofibers. Notched beam flexure tests (SEVNB) of neat and each nanofiller reinforced samples suggest that, while baseline KIC of neat geopolymer improved with heat treatment, nanofibers with high interfacial bond strength showed maximum capability in further improving KIC. Among those nanofibers, 2 vol% Silicon Carbide Whisker (SCW) showed the largest improvement in KIC of geopolymer, which is ~164%. After heat treatment at 650 °C, SCW reinforcement was also found to be effective, with only ~28% lower than the reinforcing performance at 250 °C, while the performance of Alumina Nanofiber reinforced geopolymer notably reduced. SEM and EDS analysis suggested that the inhomogeneity in neat geopolymer and length of nanofibers control the reinforcing capability as well as crack propagation resistance of geopolymer. For instance, minimum length of nanofillers to toughen this geopolymer at 250 °C was required as ~2 μm. The results further suggested that the sample failure occurred due to the dominance of tensile failure of nanofibers over the interfacial separation.  相似文献   

13.
Thermal and photochemical free radical reaction products of C60 with polymethyl methacrylate (PMMA) and polystyrene (PS) in orthodichlorobenzene solution were detected by EPR (electron paramagnetic resonance). Thermal radicals (<100°C) of C60/PMMA and C60/PS samples gave single line first-derivative EPR spectra withg=2.0029. Ultraviolet photolysis of a C60/PMMA solid phase sample gave two radical species; whereas, photolysis of a C60/PS solid phase sample gave only one free radical. EPR signals were also recorded for UV and thermal C60 reaction with free radical initiator, azobis(isobutyronitrile).  相似文献   

14.
In this study, the electrical, optical and mechanical properties of polystyrene (PS) thin films added graphene nanoplatelet (GNP) have been investigated. Surface conductivity (σ), absorbance intensity (A) and tensile modulus of these composite films have increased with increasing the content of GNP in the composite. The increase in the electrical and optical properties of the PS/GNP composite films has been interpreted by site and classical percolation theory, respectively. The electrical and the optical percolation thresholds of PS/GNP composite films were determined as Rσ?=?23.0?wt.% and Rop?=?13.0?wt.%, respectively. While the conductivity results have been attributed to the classical percolation theory, the optical results have attributed to the site percolation theory. The electrical (βσ) and the optical (βop) critical exponents were calculated as 2.54 and 0.40, respectively. The tensile modulus and the tensile strength of the PS/GNP composites increased with the increasing of GNP content in the PS. But, the toughness of the composites fluctuated with GNP addition.  相似文献   

15.
Abstract

The interphase boundary of incompatible polymer blends such as poly(methyl methacrylate) (PMMA)/natural rubber (NR) and polystyrene (PS)/NR, and of compatible blends such as PMMA/NR/epoxidized NR (ENR) and PS/NR/styrene–butadiene–styrene (SBS) block copolymer, where ENR and SBS were used as compatibilizers, was studied by means of microindentation hardness (H) and microscopy. Cast films of neat PMMA and PS, and blended films of PMMA/NR, PS/NR, PMMA/NR/ENR, and PS/NR/SBS were prepared by the solution method using a common solvent (toluene). Hardness values of 178 and 173 MPa were obtained on the surfaces of the neat PMMA and PS, respectively. After the inclusion of soft phases, the binary (incompatible) and the ternary (compatible) blend surfaces show markedly lower H‐values. Scanning electron and optical microscopy reveal a clear difference at the phase boundary of the surface of compatible (smooth boundary) and incompatible (sharp boundary) blends. The compatibilized blends were characterized by using microhardness measurements, as having the thinnest phase boundary (~30 µm), while incompatible blends were shown to present a boundary of about 60 µm. The hardness values indicate that the compatibilizer is smoothly distributed across the interface between the two blend components. Results highlight that the microindentation technique, in combination with microscopic observations, is a sensitive tool for studying the breadth and quality of the interphase boundary in non‐ or compatibilized polymer blends and other inhomogeneous materials.  相似文献   

16.
Composites, containing different concentrations of palladium (II) acetylacetonate in polymethyl methacrylate (PMMA) matrix were prepared by vigorous mixing. PMMA was prepared by solution polymerization technique. The composites were irradiated with a 120 MeV Ni10+ beam at two different fluences of 1×1011 and 5×1012 ions/cm2 to study ion-induced effects on their dielectric, structural properties and surface morphology. AC electrical properties of these samples were studied in the frequency range 100 Hz to 10 MHz. The dielectric permittivity/loss shows frequency dependent behavior and it obeys the universal law of dielectric (i.e.f n?1) for pristine and irradiated samples at high frequency. The crystalline size and crystallinity of the composites were studied by X-ray diffraction analysis. Decrease in peak intensity after irradiation signifies the amorphization which is also responsible for decrease in T g as obtained by means of differential scanning calorimetry measurement. Fourier transform infrared spectra also support this result. Surface roughness increases upon irradiation as observed from scanning electron microscopy.  相似文献   

17.
In recent years, nickel based nanomaterials have attracted much attraction owing to their low cost and the unique catalytic properties in some special reactions. Uniform diameter polymethyl methacrylate (PMMA) nanofibers and β-cyclodextrin (β-CD) functionalized PMMA nanofibers were obtained through electrospinning. Nickel-loaded nanofibers (Ni/β-CD/PMMA) were prepared by a reductive impregnation method. With increasing addition of β-CD, the diameter of the composite nanofibers became larger. The images by scanning electron microscopy (SEM) suggested that the morphology for different β-CD loadings of the β-CD/PMMA nanofibers was uniform until the loading of β-CD was 8 wt.%; the morphology of the nanofibers then became nonuniform and nodes appeared. Energy dispersive spectrometry (EDS) revealed that the Ni nanoparticles were loaded on the surface of nanofibers successfully.  相似文献   

18.
A core–shell composite has been synthesized through in situ polymerization in emulsion with an average of 205 nm of diameter. Each composite consists of a graphene oxide (GO) core and a poly(methyl methacrylate/butyl acrylate) shell. The latex is homogeneous without any aggregation after stability testing in normal temperature for 100 d and can be applied as an ideal conductive adhesive whose glass transition temperature (T g) is under ?30 °C and lucid conductive film whose T g is above 17 °C. There exists half core–shell structure in the composite with part of GO exposed which contributes to the electrical conductivity of film formed by composite. The electrical conductivity of the composite is sensitive to humidity, increasing from 0.233 to 0.357 S m?1, while the related humidity ranges from 0% to 60%. The flexible aliphatic shell established by polyacrylate chains with nanolevel of interspaces makes it easy for hydrone to move in and interact with the oxygen groups on the chains, and then the interaction enhances the difficulty for hydrone to move out, on account of which film formed by core–shell composite can hold hydrone and exhibit advanced electrical conductivity in high humidity atmosphere.  相似文献   

19.
A hierarchically nanospherical α-Fe2O3/graphene composite with a homogeneous mono-pore size of 4 nm has been prepared using a hydrothermal method. The composite showed an extremely high rate performance and good cycling stability when applied as an anode material for lithium-ion batteries owing to its unique three dimensional architecture. A specific capacity of 110 mAh/g was obtained at an extremely high current rate of 40 A/g and recover to 830 mAh/g at 0.5 A/g after 60 cycles. After 250 cycles at 2 A/g, the composite electrode exhibited a capacity of 630 mAh/g with a columbic efficiency of 99.5 %.  相似文献   

20.
In this study, we characterized the mechanical properties of fullerence (C60) epoxy nanocomposites at various weight fractions of fullerene additives in the epoxy matrix. The mechanical properties measured were the Young’s modulus, ultimate tensile strength, fracture toughness, fracture energy, and the material’s resistance to fatigue crack propagation. All of the above properties of the epoxy polymer were significantly enhanced by the fullerene additives at relatively low nanofiller loading fractions (~0.1 to 1% of the epoxy matrix weight). By contrast, other forms of nanoparticle fillers such as silica, alumina, and titania nanoparticles require up to an order of magnitude higher weight fraction to achieve comparable enhancement in properties.  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