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 31 毫秒
1.
氧化石墨烯(GO)具有较高的比表面积,层间距大,表面拥有丰富的官能团,可以很好地分散到聚合物中,但GO导电性差。研究对GO进行还原和表面修饰,以改善石墨烯和HDPE的相容性。采用熔融混炼法制备了HDPE/石墨烯复合材料,结合力学性能、导电性能、微观结构测试,考察不同HDPE/石墨烯复合材料的导电阈值,分析影响复合材料导电性的因素,进而得出较优化的制备工艺。研究发现石墨烯添加量为7.5%时,导电通路开始形成,当石墨烯含量达到7.5%时,拉伸强度提升22.14%,拉伸模量提升21.19%。  相似文献   

2.
The aim of this study was to investigate the performance of UHMWPE/HDPE-reinforced kenaf, basalt and hybrid kenaf/basalt composites. Mechanical testing of these samples was carried out such as tensile, flexural (three-point bending) and an impact test (Charpy). Pure resin (UHMWPE/HDPE) samples were tested and compare with reinforced 10% weight fraction of kenaf, basalt and hybrid kenaf/basalt samples to identifying their contribution and potential in this new composite material. UHMWPE/ HDPE sample was produced in constant ratio 60:40 respectively via extrusion process. Basalt reinforced UHMWPE/HDPE generates the highest elastic modulus result compared to kenaf and hybrid kenaf/basalt as a reinforcement material. The tensile results of kenaf reinforcement UHMWPE/HDPE samples are significantly higher (20%) than pure blend resin, which is an indication for good performance of kenaf, basalt and hybrid kenaf/basalt to be used in UHMWPE/HDPE-blend polymers. The flexural and Charpy strengths show the drawback results, where performance of polymer is reduced 5% with the absence of kenaf. It can be concluded that kenaf, basalt and hybrid kenaf/basalt fiber successfully increase the UHMWPE/HDPE blends performance especially under tensile loading.  相似文献   

3.
Scalable and ease fabrication of high-performance graphene reinforced polyamide 66 (PA66) nanocomposites by melt-mixing were achieved by selecting ideal graphene reinforcement having high C/O ratio. In this study, single-layer amine functionalized reduced graphene oxide and multi-layer thermally exfoliated graphene oxide (TEGO) were used to investigate the influence of surface chemistry and dispersion state on crystallization behaviors, mechanical, and thermal properties of graphene reinforced PA66 nanocomposites. Both types of graphenes acted as nucleating agent but TEGO showed the better performance due to its intercalated structure formation mechanism and efficient viscous flow during melting. Mechanical results indicated that 0.5 wt% TEGO based PA66 nanocomposite showed the highest tensile properties by increasing tensile modulus and tensile strength up to 45% and 16.1%, respectively. In addition, TEGO reinforced nanocomposites showed more stable viscoelastic behavior by reaching a plateau at high temperatures and restraining long-range motion of polymer chains.  相似文献   

4.
Mechanical properties of high density polyethylene (HDPE) extended to draw ratios in the 20–40 range have been determined and compared with corresponding properties of the polymers containing particulates including rutile, carbon black, iron oxide, and mica. Shrinkage of drawn structures was studied to temperatures near the fusion of the polymer host. The degree of interaction at polymer/additive interfaces was varied by surface coating certain of the solids with standard coupling agents. Solids were found to increase tensile moduli and to decrease shrinkage, particularly at higher exposure temperatures. The magnitude of changes due to the presence of solids was shown to depend on the apparent interaction at contacts between host and additive. In a dispersion–force matrix, like HDPE, benefits were optimized when the particulates were amphoteric or neutral, rather than having pronounced acid or base interaction potentials.  相似文献   

5.
In this article, a comprehensive review is presented regarding structure, synthesis, and properties of nanofillers such as graphene oxide, nanobifiller of graphene oxide, and their polymeric nanocomposite. The information about hybrid properties and synthesis of graphene oxide–carbon nanotube, graphene oxide–montmorillonite, and graphene oxide–nanodiamond is presented. Use of nanobifiller in polymer/graphene oxide–carbon nanotube, polymer/graphene oxide–montmorillonite, and polymer/graphene oxide–nanodiamond composites was summarized. Area of polymer and graphene oxide-based nanobifiller composites is less studied in literature. Therefore, nanobifiller technology limitations and research challenges must be focused. Polymer/graphene oxide nanobifiller composites have a wide range of unexplored potential in technological areas such as automobile, aerospace, energy, and medical industries.  相似文献   

6.
The purpose of the paper is to demonstrate the effectiveness of high-aspect ratio electrochemically exfoliated graphene (EEG) as a filler in high-density polyethylene (HDPE); we use an industrially viable polymer processing technique (melt blending with melt recirculation) to ensure excellent dispersion and reinforcement at low loadings. The effects of nanofiller loading were evaluated for two different HDPE grades with two different melt flow indices (MFI) based on crystallization, tensile, and rheological properties. The findings indicate improvements in mechanical properties (tensile modulus and tensile strength) for all HDPE/EEG nanocomposite samples; however, the reinforcement was more pronounced at 0.2 wt% loading, indicating a transition from excellent dispersion at lower loadings to aggregated at higher loadings. The low and high MFI HDPE/EEG nanocomposites at 0.2 wt% EEG loading displayed an improvement of 31% and 40% in tensile modulus and 19% and 33% in tensile strength, respectively. The improved mechanical response with higher MFI nanocomposites is likely due to enhanced dispersion associated with the lower melt viscosity. Similarly, the rheological results also showed maximum increase in storage and loss modulus at a loading of 0.2 wt% EEG. In conclusion, EEG can be an effective filler if proper dispersion is achieved, which is challenging at high loadings.  相似文献   

7.
The high consumption rates of polymers generate large amounts of wastes imposing long-term adverse effects on the environment combined with a significant carbon footprint. So the appeal for a circular economy is becoming loud enough to take actions. Despite increasing interests for polymer recycling, some reserves about their mechanical performances, especially long-term properties such as fatigue resistance, are barriers to introduce more recycled polymers back into production lines. In this study, a comparison between the fatigue resistance of virgin and recycled high-/low-density polyethylene (H/LDPE) is made to provide more quantitative information to address these concerns. Although some recycled polymers (HDPE) show similar tensile properties compared to virgin ones, significant differences can be observed in their fatigue lifetime. So tensile testing alone is not sufficient to provide a complete information about the overall properties of recycled polymers. Our results show that recycling polymers does not necessarily result in reduced fatigue resistance.  相似文献   

8.
In this work, sisal nanowhiskers (SNWs) extracted from sisal fibers were used to reinforce high‐density polyethylene (HDPE) and low‐density polyethylene (LDPE). The nanocomposites were prepared by solution casting from toluene and melt mixing, both followed by melt pressing. In the case of melt mixing, the surfaces of the SNW were also chemically modified with 1 phr of vinyl triethoxy silane to improve their dispersibility and compatibility with the matrices. The SNW had an average length of 197 nm and diameter of 12 nm, and a crystallinity index of 89%. Fourier transform infrared confirmed the surface chemical modification of the SNW. The whiskers were fairly well dispersed in the matrices, regardless of the treatment or preparation method. The presence of whiskers, as well as nanocomposite preparation method, had an observable influence on the storage modulus of LDPE, but very little influence on that of HDPE. There was, however, no significant influence on the degradation behavior of both polymers. The crystallization behavior of the polymers was found to strongly depend on their morphologies. The melting and crystallization behavior of the LDPE nanocomposites were almost unchanged, while an increase in crystallinity was observed for all the HDPE nanocomposites. The tensile properties depended on the type of polymer, the treatment, and the preparation method. Generally there was an improvement in tensile modulus, and a decrease in elongation at break, but the stress at break only improved for the HDPE nanocomposites. POLYM. COMPOS., 35:2221–2233, 2014. © 2014 Society of Plastics Engineers  相似文献   

9.
Economically viable high‐density polyethylene (HDPE)/graphene nanocomposites were produced using mass produced graphene powder and an industrial twin‐screw melt‐compounding machine. Rheological and electrical properties were investigated and scanning electron microscopy was carried out to investigate graphene dispersion and its network formation in the matrix. Mechanical properties of the nanocomposites were evaluated using tensile, flexural and impact tests. Differential scanning calorimetry analysis indicated that the crystalline structure of the polymer might be affected by high loadings of graphene. SEM evaluation revealed reasonable graphene dispersion in the matrix. In addition, the amount of graphene required to form a percolated network was similar for both rheological and electrical networks. The nanocomposites exhibited a significant increase in Young's and flexural moduli without a notable reduction in impact strength up to 14 wt% graphene loading. In these experiments, compounding graphene powder with HDPE produced a clear and distinct improvement in mechanical properties at an industrially suitable low cost. POLYM. ENG. SCI., 59:675–682, 2019. © 2018 Society of Plastics Engineers  相似文献   

10.
Homogeneous dispersion and strong filler–matrix interfacial interactions were vital factors for graphene for enhancing the properties of polymer composites. To improve the dispersion of graphene in the polymer matrix and enhance the interfacial interactions, graphene oxide (GO), as an important precursor of graphene, was functionalized with amine‐terminated poly(ethylene glycol) (PEG–NH2) to prepare GO–poly(ethylene glycol) (PEG). Then, GO–PEG was further reduced to prepare modified reduced graphene oxide (rGO)–PEG with N2H4·H2O. The success of the modification was confirmed by Fourier transform infrared spectroscopy, thermogravimetric analysis, and Raman spectroscopy. Different loadings of rGO–PEG were introduced into polyimide (PI) to produce composites via in situ polymerization and a thermal reduction process. The modification of PEG–NH2 on the surface of rGO inhibited its reaggregation and improved the filler–matrix interfacial interactions. The properties of the composites were enhanced by the incorporation of rGO–PEG. With the addition of 1.0 wt % rGO–PEG, the tensile strength of PI increased by 81.5%, and the electrical conductivity increased by eight orders of magnitude. This significant improvement was attributed to the homogeneous dispersion of rGO–PEG and its strong filler–matrix interfacial interactions. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134, 45119.  相似文献   

11.
《Polymer Composites》2017,38(10):2272-2282
For the purpose of the development of poly(lactic acid)/graphene oxide composites with improved tensile properties, a stearic acid compatibilizer was used to enhance the compatibility of the graphene oxide sheets with the poly(lactic acid) polymer matrix. Graphene oxide was modified with stearic acid at different mass ratios of 1:1, 1:3, and 1:5 prior to forming the composites with poly(lactic acid). Characterization showed positive effects of stearic acid attached to GO in every mass ratio and also enhanced compatibility with the poly(lactic acid) matrix. Stearic acid could strengthen the interfacial interactions between the flat graphene oxide sheets and the poly(lactic acid) matrix resulting in improved tensile strength. The tensile strength of the poly(lactic acid)/graphene oxide/stearic acid composite with a mass ratio of graphene oxide and stearic acid 1:1 increased by 32% compared to poly(lactic acid) alone. Based on these results, the graphene oxide/stearic acid composites show potential for use as nanosheet fillers for tensile strength enhancement in poly(lactic acid). POLYM. COMPOS., 38:2272–2282, 2017. © 2015 Society of Plastics Engineers  相似文献   

12.
High‐density polyethylene (HDPE)–wood composite samples were prepared using a twin‐screw extruder. Improved filler–filler interaction was achieved by increasing the wood content, whereas improved polymer–filler interaction was obtained by adding the compatibilizer and increasing the melt index of HDPE, respectively. Then, effects of filler–filler and polymer–filler interactions on dynamic rheological and mechanical properties of the composites were investigated. The results demonstrated that enhanced filler–filler interaction induced the agglomeration of wood particles, which increased the storage modulus and complex viscosity of composites and decreased their tensile strength, elongation at break, and notched impact strength because of the stress concentration. Stronger polymer–filler interaction resulted in higher storage modulus and complex viscosity and increased the tensile and impact strengths due to good stress transfer. The main reasons for the results were analyzed. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

13.
Blends of three polyethylene (PE) samples (two HDPE grades and LLDPE) with an experimental sample of a semiflexible liquid crystalline polymer (SBH 1:1:2 by Eniricerche) have been prepared in a Brabender compounder. The processing-aid effect of the LCP has been demonstrated by the decreased energy required for extruding the blends, as compared to that needed for neat PE. The thermal properties, as studied by differential scanning calorimetry (DSC), have shown that the two components of the blends are immiscible. However, the dispersed SBH phase has been found to act as a nucleating agent for the crystallization of LLDPE, whereas no such effect was observed for HDPE. This has been taken as an indication that the phase interactions of SBH with LLDPE are more pronounced than with HDPE. The morphological study of the blends, done by scanning electron microscopy (SEM), has confirmed this conclusion. In fact, the SBH particles show a much better dispersion and a narrower size distribution in the LLDPE/SBH blends. The mechanical properties of the blends have been studied on compression-molded specimens. The results indicate that the reinforcing effect of SBH is practically none for both HDPE grades. In fact, the elongation at break decreases to very low values, and the tensile strength is also reduced, when the LCP concentration increases beyond 5–10%, whereas the tensile modulus does not vary appreciably, over the whole (0–20%) LCP range investigated. On the contrary, the tensile modulus of the LLDPE/SBH blends increases up to ca. 50%, and the elongation at break decreases more smoothly, on increasing the SBH content up to 20%. © 1995 John Wiley & Sons, Inc.  相似文献   

14.
《Polymer Composites》2017,38(10):2098-2105
High‐density polyethylene (HDPE)/polypropylene (PP) composites with graphenes were prepared by melt‐compounding method. Graphene sheets were prepared through thermally reduced graphene oxide (TRG) and surfactant exfoliated graphene (SEG), respectively. Structural characterization showed that the TRG sheets exhibited a few‐layers composition with more defects compared to the SEG sheets. Morphological observations of the composites demonstrated that the graphene was preferentially dispersed in the HDPE phase and the addition of graphene (TRG and SEG) influenced the phase structure of the HDPE/PP composites. The distribution of the TRG sheets in the HDPE phase was better than the SEG sheets, and the obtained HDPE/PP composites exhibited a low electrical percolation threshold with the highly dispersed graphene. The TRG/HDPE/PP composite showed a low electrical percolation threshold of 3 wt% (1.25 vol%). For the SEG/HDPE/PP system, the percolation threshold was 7 wt% (2.98 vol%). Differences in the behavior of the two graphene components (TRG and SEG) in the HDPE/PP composites influenced the formation of percolation networks and electrical properties. POLYM. COMPOS., 38:2098–2105, 2017. © 2015 Society of Plastics Engineers  相似文献   

15.
《Polymer Composites》2017,38(10):2321-2331
In this article, reduced graphene oxide/polyimide resin composites which exhibited enhancements in mechanical properties were successfully fabricated by hot‐pressing, and reduced graphene oxide nanosheets were synthesized by thermal reduced method, which can readily mix with PI powders in aqueous solution by sonication process. The chemical structures of rGO were carefully characterized by X‐ray diffraction, Fourier transfer infrared spectroscopy and X‐ray photoelectron spectroscopy. The field emission scanning electron microscopy observations showed that the rGO displayed excellent dispersibility and compatibility with the PI matrix. The mechanical analysis indicated that the tensile and flexural strength values of the rGO/PI resin composite with 1.5 wt% rGO loading reached 80.7 and 133.3 MPa, respectively. Compared with pure PI, the optimized rGO/PI resin composite exhibited an enhancement of 30% in tensile strength, 19% in flexural strength and 27% in impact strength, due to the fine dispersion of high specific surface area of graphene nanosheets and the good adhesion between the rGO and the matrix. In addition, thermogravimetric analysis, dynamic mechanical analysis, and dielectric properties were also investigated. POLYM. COMPOS., 38:2321–2331, 2017. © 2015 Society of Plastics Engineers  相似文献   

16.
In this study, physical properties of high‐density polyethylene (HDPE) films, blended and reinforced with small amounts of liquid crystal polymer (LCP) and organoclay (org‐clay), were investigated by employing different characterization tools such as X‐ray diffractometer, scanning electron microscope, differential scanning calorimetry, rotational rheometer, dynamic mechanical analysis, and gas permeability measurements. Viscoelastic properties of samples were quantified by applying several test procedures in melt and solid‐state dynamic measurements. It was found that rigid LCP droplets were dispersed well into HDPE matrix and improved the melt elasticity and creep resistance of HDPE. Compatibilizer and org‐clay loading into HDPE/LCP blends yielded formation of smaller LCP droplets and reduced mean relaxation time and shear modulus values, compared to unloaded counterparts. It was observed that org‐clay stacks were dispersed into HDPE phase due to using of maleic anhydride grafted polyethylene as compatibilizer and HDPE/LCP/org‐clay ternary nanocomposites exhibited intercalated microstructure. Solid‐state uniaxial tensile creep behaviors of films were modeled with the Findley power‐law and four‐element Burger models. HDPE/LCP/org‐clay (90/10/5) ternary nanocomposite film exhibited better gas barrier performance than HDPE by decreasing its permeability by 50%. POLYM. ENG. SCI., 59:1344–1353 2019. © 2019 Society of Plastics Engineers  相似文献   

17.
We have investigated melt blending approaches to interfacial localization of few-layer graphene in cocontinuous polymer blends with polyethylene as one of the components. When linear low-density polyethylene (LLDPE)/polypropylene (PP) or high-density polyethylene (HDPE)/polylactic acid (PLA) and graphene were mixed all together, graphene preferred polyethylene over PP or PLA. When PP and graphene were premixed and blended with polyethylene, some graphene was trapped at the blend interface but not enough to cover the large interfacial area. In contrast, an ultralow electrical percolation was achieved (< 0.1 vol%) in HDPE/PLA blend due to smaller interfacial area. In another approach, polystyrene was added as a tertiary minor component to HDPE/PLA blends. This continuous interfacial layer containing graphene led to a low electrical percolation threshold (< 0.2 vol%). From these investigations, we suggest general ways to reduce a percolation threshold by kinetic control of the morphology of cocontinuous polymer blends.  相似文献   

18.
Hybrid nanocomposites have the unique ability of enhancing material properties due to the existing synergistic effect of the fillers. In this study, the authors report such an eco-friendly hybrid nanocomposite comprising of polyaniline and reduced graphene oxide in polycaprolactone. The conducting polyaniline improved the processability of polycaprolactone, and the final composites were prepared by incorporating graphene oxide reduced at 200 and 600°C temperatures to the polycaprolactone–polyaniline blend. Polyaniline, polyaniline/reduced graphene oxide200, and polyaniline/reduced graphene oxide600 imparted good electrical conductivity to polycaprolactone, and the fabricated flexible polycaprolactone–polyaniline/reduced graphene oxide nanocomposites exhibited good mechanical property, increased thermal stability, and excellent electromagnetic interference shielding up to 42 dB at 13 GHz.  相似文献   

19.
近年来,氧化石墨烯(GO)复合吸附材料在水处理方面得到了广泛研究.由GO和天然高分子组成的复合材料具有吸附能力强、机械稳定性好等优点,且表面改性的GO可以进一步提高其吸附稳定性和吸附能力.介绍了GO/天然高分子复合吸附材料的制备方法;总结了GO表面官能团,如巯基、氨基、羧基等对吸附材料吸附性能的影响;综述了GO与壳聚糖、纤维素、海藻酸钠、木质素等天然高分子形成的复合吸附材料对水中染料和重金属离子处理的应用;最后对GO/天然高分子复合吸附材料在实际应用中的发展方向提出了展望.  相似文献   

20.
The graphene oxide (GO) was prepared by sonication‐induced exfoliation from graphite oxide, which was produced by oxidation from graphite flakes with a modified Hummer's method. The GO was then treated by hydrazine to obtain reduced graphene oxide (rGO). On the basis of the characterization results, the GO was successfully reduced to rGO. Acrylonitrile–butadiene rubber (NBR)–GO and NBR–rGO composites were prepared via a solution‐mixing method, and their various physical properties were investigated. The NBR–rGO nanocomposite demonstrated a higher curing efficiency and a change in torque compared to the gum and NBR–GO compounds. This agreed well with the crosslinking density measured by swelling. The results manifested in the high hardness (Shore A) and high tensile modulus of the NBR–rGO compounds. For instance, the tensile modulus at a 0.1‐phr rGO loading greatly increased above 83, 114, and 116% at strain levels of 50, 100, and 200%, respectively, compared to the 0.1‐phr GO loaded sample. The observed enhancement was highly attributed to a homogeneous dispersion of rGO within the NBR matrix; this was confirmed by scanning electron microscopy and transmission electron microscopy analysis. However, in view of the high ultimate tensile strength, the NBR–GO compounds exhibited an advantage; this was presumably due to strong hydrogen bonding or polar–polar interactions between the NBR and GO sheets. This interfacial interaction between GO and NBR was supported by the marginal increase in the glass‐transition temperatures of the NBR compounds containing fillers. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42457.  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