DOI:
10.1039/C6RA14560D
(Paper)
RSC Adv., 2016,
6, 87341-87352
Effects of 2-D transition metal carbide Ti2CTx on properties of epoxy composites
Received
5th Junes 2016
, Accepted 26th August 2016
First published on 29th August 2016
Abstract
Here, 2-D nanostructured Ti2CTx was successfully synthesized by selectively exfoliating the Ti2AlC MAX phase via a liquid etching technique. Investigated as a promising and functional reinforcement additive, 2-D Ti2CTx nanosheets were dispersed into an epoxy matrix evenly at different mass percentages to obtain self-lubricating and anti-friction Ti2CTx/epoxy resin (EP) nanocomposites with high toughness and low creep strain by in situ intercalative polymerization. The phase structure and morphology of Ti2CTx and Ti2CTx/EP nanocomposites were investigated by X-ray diffraction (XRD) and scanning electron microscopy (SEM). Subsequently, this work studied the effect of Ti2CTx on the EP matrix in terms of fracture surface morphology, glass transition temperature (Tg), storage modulus (Es), thermal stability, mechanical properties and dry sliding anti-friction properties. The results show that Ti2CTx interlayers were intercalated and exfoliated with EP molecular chains and intertwined with each other, which produced a greater interfacial area between the inorganic additive and the polymer matrix, resulting in the formation of Ti2CTx/EP nanocomposites without adverse structural defects. With the addition of an appropriate mass fraction of Ti2CTx, the fracture toughness and flexural strength of EP composites reached 17.8 kJ m−2 and 98 MPa, which represented increases of 76% and 66%, respectively, compared with those of pure EP. In addition, increases in Tg and Es of 15% and 38% were achieved, respectively, relative to those of pure EP when the concentration of Ti2CTx was 2.0 wt%. Moreover, tribological analysis indicated that there was a great improvement in anti-friction properties and the morphology of worn surfaces of the nanocomposites appeared to be smoother than that of pure EP, which implied the transformation of fatigue and abrasion wear into adhesion wear. This work is of great significance in that novel 2-D transition metal carbide crystals were employed to fill and modify a thermoset polymer resin and enhance its toughness and strength and, as a self-lubricating additive, enabled a great breakthrough in the anti-friction properties of polymer-matrix composites.
1. Introduction
Epoxy resin (EP), as a high-performance thermoset synthetic polymer resin, possesses prominent advantages such as a low shrinkage rate, high adhesive strength, corrosion resistance, strong dielectric properties, environmental stability, relatively low cost, etc., and has been used in the design of intricate structural parts employed in aerospace and automobile components and electrical insulation.1,2 In actual applications, higher requirements have been proposed for polymer materials that took the place of wood and steel. However, its high brittleness, low toughness and resistance to thermal shock, and poor friction and wear properties have restricted the development of EP in applications, especially in anti-friction and high-frequency vibrating parts in machinery. Therefore, it is a difficult problem to increase the toughness and enhance the creep resistance and anti-friction properties of EP by physical or chemical methods.
Currently, the improvement of polymer performance via adding fillers to synthesize polymer-matrix composites has become a research focus.3 Taking into consideration the excellent properties of inorganic nanomaterials, they are widely used as fillers to improve the properties of polymers. Mayoral et al. synthesized graphene nanoplatelet (GNP)/polyamide 6 composites by melt processing, which displayed a 420% increase in Young's modulus at a 20 wt% loading of nanofiller.4 Chemical oxidative polymerization has been employed to synthesize magnetite/SPAN conductive polymer-based nanocomposites, which have satisfactory conductivity.5,6 Choi et al. employed a sol–gel method to prepare graphene (FGS)/waterborne polyurethane nanocomposites, in which the finer dispersion of FGS improved the conductivity and thermal resistance of polyurethane and acted as a reinforcing filler to increase the tensile strength and modulus.7 Ultrasonic irradiation polymerization was employed to synthesize nano-SiO2/poly(3-aminophenylboronic acid) (PAPBA) composites with a core–shell morphology, which exhibited a high conductivity of 0.2 S cm−1.8 In addition, microemulsion polymerization is also a vital way of preparing organic–inorganic nanocomposites. Reddy et al.9 reported that microemulsion polymerization was employed to coat MWCNTs with polyaniline (PANI) nanospheres and proved that a MWCNTs/PANI nanocomposite without a surfactant had higher conductivity than a composite synthesized by in situ chemical polymerization. Besides, utilizing in situ chemical oxidative graft polymerization,10 a hybrid MWCNTs-f-PANI-noble metal nanocomposite was synthesized and displayed significantly higher conductivity than those of CNTs-f-PANI and pure PANI.
However, the synthesis methods of epoxy-based nanocomposites have mainly employed direct mixing and intercalative polymerization because of the fluidity of epoxy monomers. Direct mixing is one of the easiest methods of synthesizing nanoparticles/EP composites11 via sonication and mechanical stirring, which disperses surface-treated nanoparticles into an EP matrix and minimizes agglomeration. Owing to the unique microscopic morphology of two-dimensional (2-D) nanomaterials (e.g., graphene,12–14 nanoclay,15 ceramics,2 MoS2,16,17 BN,18 etc.), they can be exfoliated into monolithic nanolayers and filled into an EP matrix to manufacture high-performance composites by in situ intercalative polymerization. Via in situ intercalative polymerization, a 2-D layer nanostructure bonded by van der Waals bonds is intercalated by epoxy molecular chains, which results in an increase in the interlayer distance and separation into strips after curing.3,19–21 Moreover, the exfoliated nanosheets embedded in the epoxy matrix could produce a larger interfacial surface area between the inorganic additive and the polymer matrix. Moreover, because in situ intercalation achieves exfoliation, 2-D lamellar nanosheets play a better role in interfacial interactions with an epoxy matrix.15,22
Recently, new types of 2-D transition metal carbides and carbonitrides called MXenes, which possess graphene-like 2-D morphology23,24 with a general formula of Mn+1XnTx, where M is an early transition metal, X is C and/or N, n is 1–3, T is a surface terminating functionality (O, F and/or OH), and x represents the number or type of functional groups,23,25 were fabricated by selective etching of the A element (an A-group element, mostly group IIIA or IVA) layers from MAX phases, which are a large family of ternary layered carbide and nitride ceramics,26 and attracted much attention and triggered numerous applications.27–31 Owing to their novel electrical properties,32–34 excellent mechanical strength,35 adsorption properties,36 etc., MXenes have been widely used in research into new energy materials,37–41 lubricant additives,42,43 and polymer-reinforced materials.44,45
To our knowledge, hardly any literature has reported MXene/EP composites. Here, taking advantage of the layered structure and surface terminating functionalities of MXenes, Ti2CTx/EP composites were synthesized by means of in situ intercalative polymerization, as shown in Fig. 1, and the effects of the 2-D transition metal carbide Ti2CTx on the properties of epoxy-based composites were investigated in detail. Moreover, MXene nanosheets played a role as a reinforcing framework to support crosslinked EP molecular chains, as graphene46 or the 2-D MAX phase Carbide Derived Ceramic Nanostructure (CDCN)2 played the role in an EP matrix of increasing the adhesion force and acting as an interlocking chain between the nanofillers and the EP matrix in previously reported papers. A Ti2CTx/EP nanocomposite was prepared successfully, which had better mechanical, thermal, creep resistance and anti-friction properties than those of pure EP. These results show that 2-D MXene nanosheets have significant potential as reinforcing additives in polymer-based composites.
 |
| | Fig. 1 Schematic diagram of EP interlayered into Ti2CTx layers: (a) Ti2CTx was dispersed into acetone and mixed with EP, where the two phases separated without interacting; (b) following ultrasonic vibration and mixing, EP monomer was intercalated into Ti2CTx layers; (c) cured epoxy resin with extensive separation of Ti2CTx into strips. | |
2. Experimental
2.1 Materials
The MAX phase Ti2AlC, which is a precursor of the 2-D nanomaterial Ti2CTx, was synthesized by pressless sintering in our previous report.47 The pristine Ti2CTx that was utilized in this study was synthesized by selectively exfoliating A atoms from the MAX phase in a solution of lithium fluoride (LiF, 99%, Tianjin Fine Chemical Co., Ltd, China) in hydrochloric acid (HCl, AR grade). Epoxy resin (DGEBA, epoxy equivalent 184–195 g mol−1, viscosity 10–16 Pa s) was purchased from Xingchen Synthetic Material Co., Ltd (China) and 2-ethyl-4-methylimidazole (EMI-2,4, 98%, C6H10N2, MW 110.16 g mol−1) hardener and antifoaming agents (AF) were supplied by Chinese Macklin Biochemical Co., Ltd. Acetone and ethyl alcohol (AR grade, 99%) were supplied by Hongyan Reagent Company of China and distilled water (DW) was lab-synthesized.
2.2 Preparation of 2-D Ti2CTx MXene
The novel 2-D nanostructured MXene Ti2CTx was synthesized by solution etching.48 Initially, 5.0 g LiF was dispersed into 100 mL 6 M HCl solution in a PTFE container. After ultrasonic concussion for 10 min, 5.0 g Ti2AlC was weighed and added to the above solution over 5 min and ultrasonic vibration was continued for 30 min. Afterwards, the container was transferred to a magnetic stirring apparatus and stirred for 48 hours at 40 °C. The resulting suspension was subjected to ultrasonic treatment for 10 min and then centrifuged at 5000 rpm for 10 min. The supernatant was decanted and the remaining sediment was washed with DW several times until the pH value of the supernatant was close to neutral. Finally, the sediment was washed with ethyl alcohol three times and dried in a vacuum at 80 °C for 12 h.
2.3 Fabrication of Ti2CTx/EP nanocomposite
Initially, different masses (0.5 g, 1.0 g, 1.5 g and 2.0 g, respectively) of the prepared Ti2CTx were weighed, dispersed into 10 mL acetone and subjected to ultrasonic vibration for 30 min at room temperature (RT). Then EP (100 g) and AF (0.25 g) were added to the above solutions with simultaneous mechanical stirring at a speed of 1500 rpm and ultrasonic treatment for 1 hour at RT and then heated to 50 °C with continued stirring at the same speed, again with ultrasonic treatment, for 30 min. An intermediate degassing procedure was carried out at 60 °C for 5 hours to remove acetone and trapped air. Afterwards, the hardener EMI-2,4 (5 g) was added and stirring and ultrasonic concussion were continued for one hour, then the mixture was degassed by vacuum-pumping for 2 hours at RT. The resulting mixture was highly degassed, followed by being poured into a PTFE mould, which had been preheated at 80 °C in order to ensure better wetting between the mixture and the mould. Afterwards, the curing process included three stages, namely, pre-curing at 60 °C for 1 h, curing at 100 °C for 2 h and post-curing at 120 °C for 2 h. The whole process was carried out in a high-temperature drying oven.
2.4 Characterization and measurement
2.4.1 Crystal phase and microstructure observation. X-ray diffraction (XRD, SmartLab X-ray diffractometer equipped with copper target Kα radiation, λ = 1.5406 Å, Rigaku Corporation, Japan) was used to detect the crystal structures of Ti2AlC, Ti2CTx powder and Ti2CTx/EP nanocomposites. The samples were scanned over the two theta angle range (2θ) of 2–80° in steps of 0.02° and at a scanning speed of 10° per min to obtain the diffraction patterns.The morphology and microstructures of the powders and nanocomposites were examined by a field emission gun scanning electron microscope (FEG-SEM, accelerating voltage 10 kV, FEI Quanta 250, Germany). The samples were fixed on a flat metal pallet with conductive adhesive tape and the test surface was sprayed with gold for 90 seconds.
2.4.2 Dynamic mechanical and thermal analysis. Dynamic mechanical analysis (DMA) was conducted using a dynamic mechanical analyzer (DMA Q800, TA Instruments, Inc., USA) equipped with a single cantilever clamp, and the rectangular test specimens had dimensions of 35 mm (length) × 12.5 mm (width) × 2.0 mm (thickness). The test was carried out by a temperature sweep (temperature ramp from 50 °C to 220 °C at 3° per min) method at a constant frequency of 1 Hz to obtain the dynamic storage modulus and damping curves (tan delta) of the specimens. The Tg values obtained from the maximum peaks in the tan delta curves correspond to the Tg temperature.Creep and recovery tests were performed using the above-mentioned DMA with the same sample dimensions. The creep properties were characterized as a function of creep time for 20 min and recovery time for 20 min under applied stress levels of 5, 10 and 15 MPa at 75 °C.
Thermal gravimetric analysis (TGA) of samples was carried out using a Setaram Evolution 2400 thermal analyzer under an argon flow of 20 mL min−1 with a heating rate of 5 °C min−1 from room temperature to 600 °C.
2.4.3 Mechanical properties measurement. The microhardness of pure EP and Ti2CTx/EP nanocomposites was investigated using a Q/NFAK1-2000 Vickers microhardness tester (MH-5, Shanghai Hengyi Electronic Testing Equipment Co., Ltd, China) with a Vickers diamond pyramidal indenter having a square base and a pyramidal angle of 136°. The specimens were subjected to a load of 2 N with a dwell time of 20 s and received indentations. The microhardness of each specimen was measured five times and the results were averaged.The impact strength was measured by a Charpy impact strength tester (XJ-50Z, Combination of Impact Testing Machine, China), which was equipped with a pendulum with an impact energy of 7.5 J and a blade angle of 30°, which fell from an elevation of 160° with an impact velocity of 3.8 m s−1 to fracture the specimens to obtain the value of the impact energy. The dimensions of the specimens were 50 mm (length) × 6.0 mm (width) × 4.0 mm (thickness). The results for the value of the impact energy were obtained by measuring three times and averaging.
Flexural strength tests on the composites were performed with a WDW-20 Electronic Universal Testing Machine (China) in three-point bending test mode, which had a load cell capacity of 10 kN. The elongated test specimens were prepared with dimensions of 40 mm (length) × 10.0 mm (width) × 2.0 mm (thickness). The span distance of the three-point bending test was 32 mm. The crosshead had a speed of 2 mm min−1. The values of the flexural strength were obtained by measuring three times and then averaging.
2.4.4 Anti-friction measurements. The tribological properties of Ti2CTx/EP composites were investigated with pin-on-disk pairs provided by a multi-specimen test machine, which was consistent with our previous method reported in the literature.45 As the disk pairs, specimens with dimensions of 40 mm × 40 mm × 4.0 mm and the frictional surfaces of steel pins and disks were polished with 1200 mesh SiC metallographic sandpaper and washed with acetone. Friction testing was conducted at rotating speeds of 100 rpm (linear speed 0.15 m s−1) and 200 rpm (linear speed 0.3 m s−1) under loads of 98 N for a test duration of 30 min to investigate the effect of the addition of different amounts of Ti2CTx and the slipping velocity on the frictional coefficient of Ti2CTx/EP nanocomposites.
3. Results and discussion
3.1 Polymerization mechanism for the formation of composites
The typical polymerization mechanism of DGEBA cured by EMI-2,4 hardener is shown in Fig. 2. Firstly, the hardener acting as a catalyst gave rise to ring-opening and breakage of C–O bonds in epoxy monomer and moreover N–H bonds are broken and H is transferred from N–H to O to form the intermediate. Secondly, the tertiary amine initiated the ring-opening of EP monomer and began the polymerization reaction. On the addition of Ti2CTx, we inferred that the surface OH functional groups bonded to the epoxy groups with simultaneous ring-opening, in a similar way to an etherification reaction. In other words, the addition of Ti2CTx played the role of a bridge that promoted the curing reaction and the formation of a crosslinked network in the EP matrix. As shown by the results of FT-IR spectroscopy in Fig. 3, the fingerprint peak of epoxy groups in epoxy monomer at 915 cm−1 is significant; after curing, the disappearance of the fingerprint peak means that the initiator has broken C–O bonds in the epoxy group and induced ring-opening. Moreover, peaks due to OH group stretching vibrations appeared at 3600–3800 cm−1 for the cured EP and Ti2CTx/EP nanocomposite,2 corresponding to the ring-opening of epoxy groups and introduction of OH groups in the curing process in Fig. 2.
 |
| | Fig. 2 Reaction mechanisms and polymerization process of DGEBA cured by EMI-2,4 hardener. | |
 |
| | Fig. 3 FT-IR spectra of epoxy monomer before and after curing and Ti2CTx/EP composites with the addition of 2.0 wt% Ti2CTx. | |
3.2 Phase and morphology analysis of Ti2CTx and Ti2CTx/EP nanocomposites
Fig. 4 shows the XRD patterns of MAX phase Ti2AlC, MXene Ti2CTx, pure EP and a Ti2CTx/EP composite. Obviously, we can see that the diffraction peaks of the (002), (103), and (106) crystal planes of Ti2AlC, which corresponded to values of 2 theta of 12.9°, 39.4°, and 53.1°, respectively, disappeared after liquid etching. Simultaneously, a new diffraction peak appeared at 7.37°, which corresponded to the (002) crystal plane of Ti2CTx.49 Moreover, the XRD pattern of pure EP indicated its amorphous nature based on the broad diffraction hump at a 2 theta angle of about 17.5°.50 It can be seen in the inset of Fig. 4 that the d-spacing of Ti2CTx for (002) crystal planes increased from 11.7 Å to 14.4 Å and correspondingly the 2 theta angle shifted from 7.37° to 6.23°; in other words, the diffraction peak shifted left to a lower angle and revealed an increase in the layer spacing. In addition, the intensity of the (002) diffraction peak of Ti2CTx in the composite decreased and the full width at half maximum (FWHM) was significantly greater compared with that for Ti2CTx. The reason for this result is that chains of epoxy molecules were intercalated into the interlayer and interweaved with Ti2CTx, causing an increase in the layer spacing of Ti2CTx and even destruction of the crystal structure, as in the observation that the structures of silicate clay or graphite were intercalated by molecular chains of epoxy resin in previous reports.50,51
 |
| | Fig. 4 XRD patterns of MAX phase Ti2AlC, MXene Ti2CTx, pure EP and EP nanocomposite with a Ti2CTx content of 2.0 wt%. | |
Fig. 5 shows a schematic diagram of the preparation process of Ti2CTx and Ti2CTx/EP composites. Fig. 5(a) explains the etching of the ternary layered ceramic MAX phase Ti2AlC to give the MXene Ti2CTx at the level of atomic structure, in which the active metallic element Al in the MAX phase was exfoliated and the remaining atomic layers of transition metal and carbon were bonded by covalent bonds to form the stable 2-D transition metal carbide Ti2CTx, which is similar to treating the MAX phase with hydrofluoric acid.49 Subsequently, EP molecular chains were intercalated into Ti2CTx layers and cured, resulting in an increase in the layer spacing. Fig. 5(b) and (c) shows SEM pictures of Ti2AlC particles and 2-D layer-structure Ti2CTx, in which the thickness of a single layer is about 50 nm and the spacing between Ti2CTx layers is approximately 100 nm and molecular chains of epoxy resin were easily intercalated into the interlayers. Fig. 5(d) shows the fracture surface of the cured Ti2CTx/EP nanocomposite. It can be observed that layered Ti2CTx was exfoliated into a single layer or few layers and dispersed in the EP matrix homogeneously, which demonstrated the good compatibility of EP and Ti2CTx.
 |
| | Fig. 5 (a) Schematic diagram of preparation of Ti2CTx and Ti2CTx/EP nanocomposite; SEM of MAX phase Ti2AlC (b), which was etched to form 2-D layer-structure Ti2CTx (c) and a Ti2CTx/EP nanocomposite (d). The inset in (b), (c) and (d) is the high magnification image. | |
The fracture morphology of pure EP and Ti2CTx/EP nanocomposites is shown in Fig. 6. There only appear to be pores of a few microns in size on the relatively smooth fracture surface of pure EP (Fig. 6a), which exhibits brittle fracture characteristics. Fig. 6(b)–(d) shows the fracture surfaces of Ti2CTx/EP nanocomposites, which are rougher than that of pure EP owing to the addition of Ti2CTx. Clearly, Ti2CTx layers were intercalated and expanded (Fig. 6b), and were exfoliated into single layers and embedded in the EP matrix (Fig. 6c). Furthermore, it was demonstrated clearly that layered Ti2CTx was dispersed uniformly in the polymer matrix without agglomeration on a cross-section of the EP composite (Fig. 6c and d). In addition, the fracture surfaces of Ti2CTx/EP nanocomposites exhibited a typical flow pattern, which can be used to estimate the local direction of crack propagation, which is marked by the black arrows on the cross-sections of the composites.52,53 With an increase in the Ti2CTx content, the number of micro-cracks increased and the fracture surface displayed typical tough fracture features. A similar phenomenon has been observed in previous research.21,53,54
 |
| | Fig. 6 FE-SEM images obtained of the fracture surfaces of cured samples: (a) pure EP, (b) molecular chains of EP intercalated into Ti2CTx layers resulting in an increase in the layer spacing, (c) and (d) cured epoxy resin with Ti2CTx sheets extensively exfoliated into single layers. | |
3.3 Analysis of thermal, mechanical and anti-friction properties of composites
3.3.1 Dynamic mechanical analysis and thermal stability. The thermal stability of polymer materials plays a vital role in engineering applications. In particular, the glass transition temperature (Tg) is one of the most critical parameters, which indicates the upper temperature limit for the application. Moreover, changes in Tg values have a crucial relationship with reinforcement phases and the interfacial interaction between reinforcement materials and the epoxy resin matrix.15,55 Fig. 7(a) shows the loss factor (tan delta) of pure EP and EP composites with different mass fractions of Ti2CTx. The Tg value was determined from the temperature at the peak of the corresponding glass transition region of the tan delta curve, which increased gradually with a change in the mass concentration of Ti2CTx (Table 1). Obviously, the Tg of pure EP is 141.5 °C, whereas when the concentration of Ti2CTx increased to 2.0 wt%, the Tg of the Ti2CTx/EP nanocomposite increased by ∼20 °C and ∼15%, respectively. In other words, the addition of Ti2CTx can increase the Tg value dramatically, which can be explained by the interfacial interaction and crosslinking between Ti2CTx nanosheets and EP molecular chains. As mentioned above, EP molecular chains were grafted on the hydroxyl-functionalized surface of Ti2CTx nanosheets and in combination with the nanosheets interweaved into a three-dimensional crosslinked network structure during the curing process. When the temperature increased to the glass transition region, because the Ti2CTx surface developed a strong interaction with molecules in the matrix, stiff nanosheets hindered the vibration of molecular chains and caused an increase in the relaxation time, which led to an increase in the value of Tg.15,56–58
 |
| | Fig. 7 Results of DMA and TGA: (a) tan delta curve, from which was obtained the glass transition temperature (Tg) and (b) storage modulus (Es), (c) TG and (d) DTG of pure EP and Ti2CTx/EP nanocomposites. | |
Table 1 Detailed parameters of the results of analysis of the thermal properties of Ti2CTx/EP nanocomposites with different amounts of Ti2CTxa
| Sample |
DMA |
TGA |
| Tg (°C) |
Es (GPa) |
Tonset (°C) |
Td10 (°C) |
Tdm (°C) |
ΔHd (J g−1) |
| Tg: glass transition temperature; Es: storage modulus at 50 °C; Tonset: starting degradation temperature; Td10: decomposition temperature of 10 wt% weight loss; Tdm: fastest decomposition temperature at the peak of the DTG curve; ΔHd: degradation enthalpy per gram. |
| Pure EP |
141.5 |
1.33 |
383.6 |
401 |
438.2 |
46.9 |
| Ti2CTx/EP (0.5%) |
152.1 |
1.31 |
383.9 |
401 |
438.8 |
47.2 |
| Ti2CTx/EP (1.0%) |
155.0 |
1.82 |
373.4 |
399 |
437.6 |
40.0 |
| Ti2CTx/EP (1.5%) |
155.6 |
1.79 |
373.2 |
398 |
437.1 |
37.1 |
| Ti2CTx/EP (2.0%) |
162.1 |
1.77 |
372.7 |
395 |
435.9 |
35.6 |
The storage modulus (Es) reflects the viscoelasticity and rigidity of materials, which determine their ability to resist deformation when affected by external forces of constant frequency and temperature. Fig. 7(b) shows that as the temperature increased there was a change in curves of Es versus different contents of Ti2CTx added to the EP matrix, indicating a marked increase in Es (storage modulus at 50 °C; the values are listed in Table 1). Moreover, the viscoelastic transition temperature of EP composites increased with an increase in the content of Ti2CTx nanosheets. The main reason is probably that the reinforcing effect of Ti2CTx nanosheets and the large specific surface area and unique two-dimensional planar structure of Ti2CTx enhanced the adhesion and interlocking action between the nanofiller and the EP molecular chains and improved the interfacial interaction, resulting in a consequent increase in the toughness of Ti2CTx/EP composites (Table 2).12
Table 2 Detailed parameters of mechanical and tribological properties of Ti2CTx/EP nanocomposites with different amounts of Ti2CTx
| Sample |
Mechanical properties |
Friction coefficient |
| Microhardness (Hv) |
Flexural strength (MPa) |
Impact strength (kJ m−2) |
Sliding velocity |
| 0.15 (m s−1) |
0.3 (m s−1) |
| Pure EP |
23.8 ± 1.2 |
59 ± 1.4 |
10.1 ± 0.2 |
0.768 |
0.673 |
| Ti2CTx/EP (0.5%) |
19.6 ± 0.4 |
78 ± 1.7 |
15.2 ± 1.1 |
0.801 |
0.605 |
| Ti2CTx/EP (1.0%) |
18.2 ± 0.1 |
98 ± 2.0 |
17.8 ± 1.6 |
0.717 |
0.365 |
| Ti2CTx/EP (1.5%) |
18.6 ± 0.1 |
84.5 ± 2.4 |
14.6 ± 1.6 |
0.67 |
0.257 |
| Ti2CTx/EP (2.0%) |
20.8 ± 1.8 |
81 ± 1.1 |
13.3 ± 2.1 |
0.454 |
0.228 |
Fig. 7(c) and (d) shows the TGA and DTG curves of EP composites with different contents of Ti2CTx, and the detailed parameters are listed in Table 1. Obviously, the starting degradation temperature (Tonset), fastest decomposition temperature (Tdm), 10% weight loss temperature (Td10) and degradation enthalpy per gram (ΔHd) generally decreased as the content of Ti2CTx in the composites increased. These results can be explained because Ti2CTx nanosheets with high thermal conductivity were finely dispersed in the EP matrix and displayed a rapid thermal response;59 when the temperature rose the motion and rearrangement of EP molecular chains were hindered by the dispersed Ti2CTx nanosheets60 owing to their mutual interactions. Therefore, the addition of Ti2CTx nanosheets will possibly reduce the thermal stability of the composites to a certain extent.
3.3.2 Creep and recovery behavior. As we know, it is extremely important to improve the creep resistance of polymer materials, in particular for their long-term use under a certain load and temperature. Fig. 8 shows the creep and recovery properties of pure EP and Ti2CTx/EP composites with different contents of Ti2CTx under different applied stresses in relatively high-temperature conditions. Clearly, a higher level of creep stress resulted in larger strain in pure EP and its composites, as shown in Fig. 8(a)–(c). Without regard to the change in levels of applied stress, compared with that of pure EP, the creep strain of the composites underwent a clear reduction with an increase in Ti2CTx content. Furthermore, it is notable that the instantaneous and highly elastic deformation of Ti2CTx/EP composites underwent a greater recovery after removal of the applied stress, and the composites exhibited a lower permanent deformation than pure EP, which indicated that there was a significant improvement in creep recovery performance with the addition of Ti2CTx nanosheets to the EP matrix. Fig. 8(d) shows the creep strain and unrecovered strain of pure EP and its composites with different contents of added Ti2CTx under high levels of applied stress at 75 °C. It is worth noting that the creep strain and unrecovered strain of the composite with 2.0 wt% Ti2CTx nanosheets after a creep time of 20 min at a stress level of 15 MPa and removal of stress with a recovery time of 20 min displayed reductions of ∼20% (creep strain ∼1.1%) and ∼40% (unrecovered strain ∼0.06%) compared with those of pure EP (∼1.33% and ∼0.11%), respectively. In general, the addition of Ti2CTx nanosheets made EP display better creep resistance and lower unrecovered strain than pure EP to a great extent at different stress levels.
 |
| | Fig. 8 Creep and recovery of pure EP and its composites with different contents of Ti2CTx under different applied stresses of 5 MPa (a), 10 MPa (b), and 15 MPa (c) at 75 °C and (d) creep strain and unrecovered strain measured at 15 MPa and 75 °C. | |
The above-mentioned differences in the creep and recovery behaviour of composites with different contents of added Ti2CTx may be explained by several reasons. Firstly, the dispersion of a filler in the matrix has a major influence on the creep behavior of polymer composites. Ti2CTx particles with a layered structure were exfoliated and intercalated and distributed in the matrix evenly, establishing a close contact with EP molecular chains, and formed a larger interfacial area to improve the interfacial bonding and compatibility.61,62 Secondly, Ti2CTx nanosheets interacted with EP molecular chains and occupied part of the free volume in the EP matrix, so that the molecular segments underwent reductions in activity and mobility. It was difficult for the molecular chains to move and furthermore for the modulus and dimensional stability of the composite to be improved, resulting in decreases in creep strain and unrecovered strain.63 In addition, EP molecular chains were combined with Ti2CTx nanosheets to form a more stable three-dimensional crosslinked network framework during the epoxy curing process. As shown in Fig. 6, the fracture surfaces of Ti2CTx/EP composites are rougher than that of pure EP, which confirmed that rigid Ti2CTx nanosheets that were intertwined with the polymer matrix produced a larger interfacial area, prevented the motion of the molecular chains and dominated the processes of interlayer sliding and the tremendous deformation of the substrate under an external force, resulting in less creep deformation and a lower strain rate in the composite than in pure EP.64–66 In essence, below the glass transition temperature pure EP and its composite exhibit mainly elastic deformation and a certain extent of viscous flow under a certain level of creep stress. After the addition of Ti2CTx nanosheets to the EP matrix, the internal friction due to the motion of EP molecular segments would increase and the composites would deform and display longer relaxation times.67 Therefore, the composites exhibited less creep strain and permanent deformation after recovery behavior.
3.3.3 Analysis of mechanical properties. Surface hardness is an important parameter for measuring the application performance of materials. Fig. 9 shows a curve of Vickers microhardness values with an increase in Ti2CTx content. It can be seen that the microhardness of EP composites decreased first and then increased with an increase in Ti2CTx content. When the concentration of Ti2CTx reached 1.0 wt%, the value was reduced from 23.8 Hv to 18.2 Hv, which is about 23.5% less than that of pure EP. This trend is most probably related to the fluidity of molecular chains on the surface of the epoxy polymer. With the addition of inorganic nanoparticles, some defects were introduced into the composites and increased the free volume in the EP matrix, along with increasing the fluidity of EP molecular chains on the surface of the composite, namely, the surface hardness decreased. However, the microhardness of Ti2CTx/EP composites increased as the Ti2CTx content continued to increase. When a certain amount of Ti2CTx was added to the EP composite material, this played the role of supporting loads and enhancing its ability to resist deformation under external stress.68,69 However, the results obtained from surface indentation tests have large standard deviation errors that should be taken into consideration. Moreover, an increase in the filler content may introduce more defects, which have an adverse effect on the mechanical properties of the composites.
 |
| | Fig. 9 Vickers microhardness of pure EP and Ti2CTx/EP nanocomposites. The inset shows indentation images observed by a metallography microscope of pure EP and the composite with a Ti2CTx content of 2 wt%. | |
Its values of impact strength and flexural strength can reflect the toughness of a material indirectly. Fig. 10 shows the changing trends in the impact fracture strength and flexural strength of Ti2CTx/EP nanocomposites. As shown in Fig. 10(a), the impact strength of EP was significantly increased with the addition of Ti2CTx nanosheets. On an increase in the Ti2CTx content to 1.0 wt%, the impact strength reached a maximum value of 17.8 kJ m−2, which represented an increase of 76% compared with that of pure EP. Similarly, in Fig. 10(b) the flexural strength reached 98 MPa when the mass content of Ti2CTx was 1.0 wt%, which, compared with Ti3SiC2/EP composites with a Ti3SiC2 content of 30 wt% (ref. 2) and Ti3AlC2/EP composites with a Ti3AlC2 content of 40 wt% (ref. 70) represented an increase of about 30% and 36%, respectively. The increase in impact fracture strength and flexural strength may have the following reasons. Firstly, the dispersion of Ti2CTx nanosheets in the polymer matrix has a major effect. According to the SEM analysis in Fig. 6, Ti2CTx nanosheets were dispersed into the EP matrix evenly and the composite exhibited good interfacial bonding. When the composite was deformed under the action of an external force, the Ti2CTx nanosheets would deform. Because the stiffness of Ti2CTx nanosheets is much greater than that of the organic matrix, when Ti2CTx nanosheets undergo deformation under longitudinal tensile stress and transverse compressive stress, this will result in hysteresis relative to the substrate in the entire process of deformation.52,53 Furthermore, they absorbed more energy and increased the impact strength and enhanced the toughness of the composites. In addition, interfacial interaction is the basis of two-phase stress transfer. Therefore, the strength of interfacial bonding has a direct effect on the ability of composites to resist deformation. During the fracture process of Ti2CTx/EP nanocomposites, the external force gave rise to interfacial debonding between the inorganic nanosheets and the organic matrix. Crack propagation encountered the rigid Ti2CTx nanosheets, which deflected the developing cracks. As the cross-section morphology shows in Fig. 6, the roughness of the fracture surface increased with an increase in Ti2CTx content, and the composite had a larger total fracture surface area and dissipated more energy during the course of impact fracture compared with pure EP, exhibiting greater impact resistance and toughness.17 Besides, the Ti2CTx nanosheets promoted the generation of shear yielding under an external force in the composites. Interfacial debonding in combination with shear yielding consumed a large amount of energy during deformation, and then the composites displayed higher strength.53,71 Nevertheless, there is a limitation on the filler content. The larger the content of fillers added to the polymer matrix, the more defects will be introduced into the composites, exactly as Fig. 10(a) and (b) shows that the impact strength and flexural strength exhibited reducing trends when the content of Ti2CTx exceeded 1.0 wt%. The current result also supports the fact that the fracture toughness of nanocomposites will decrease when the nanofiller concentration exceeds a certain weight fraction and give rise to cracks that are propagated unstably, as reported in the previous literature.53,72,73 Based on the above-mentioned analysis, it should be essential to take the appropriate amount of nanofiller added to the polymer matrix into consideration to improve the material performance.
 |
| | Fig. 10 Impact strength (a) and flexural strength (b) of composites with different mass concentrations of Ti2CTx. | |
3.3.4 Analysis of anti-friction properties. As components, the anti-friction properties of polymers are essential in engineering applications. Here, we investigated the relationship between the friction coefficient, filler content and sliding velocity by using a dry sliding friction pin-disk tester. Fig. 11 shows the friction coefficients of Ti2CTx/EP composites with different Ti2CTx contents, which were measured under different friction conditions. Obviously, it can be seen that the friction coefficient is closely related to the friction velocity and the concentration of Ti2CTx. The friction coefficient of the epoxy composite with a 2.0% mass fraction of Ti2CTx decreased by more than twice that of pure EP at a sliding velocity of 0.3 m s−1. However, at a lower sliding velocity of about 0.15 m s−1, the friction coefficient of the composite with a 2.0% mass concentration of Ti2CTx decreased by 40%, which demonstrates an obvious difference from the decrease at a high sliding speed. That is to say, it is more favorable to improve the anti-friction properties of the EP composite in high-speed sliding conditions.
 |
| | Fig. 11 Friction coefficient of pure EP and its composites obtained by dry sliding wear and the relationship of the friction coefficient with different contents of Ti2CTx in nanocomposites and the sliding velocity or distance. | |
Fig. 12 shows FE-SEM micrographs of the worn surfaces of EP with different contents of Ti2CTx nanosheets. It can be seen that there was severe wear damage on the worn surface of the pure EP matrix (Fig. 12), and the dominant wear mechanisms were fatigue and fracture wear owing to the detachment of bulk materials and formation of cracks.74,75 Compared with pure EP, the worn surfaces of EP nanocomposites with an increase in the content of Ti2CTx nanosheets displayed substantive changes. The phenomenon of fracture damage was gradually reduced and the worn surface became smoother with an increase in Ti2CTx content (Fig. 12(b)–(e)), which means that the nanofiller can transfer stress effectively and prevent the initiation of cracks and further extension of fractures. From the SEM analysis of the worn surface, it can be inferred that in addition to fracture and fatigue wear, adhesion wear also contributes. During the frictional process, Ti2CTx nanosheets distributed in the composites gave rise to surface friction with the steel pins directly and generated a large amount of heat, which softened the worn surface. Furthermore, intercalated Ti2CTx moved freely and became embedded into the softened worn surface at high loads and constituted a friction transfer film at the interface, providing a certain lubricating effect.2,43,74
 |
| | Fig. 12 FE-SEM micrographs of the worn surfaces of EP with different mass concentrations of Ti2CTx: (a) pure EP, (b) 0.5 wt%, (c) 1.0 wt%, (d) 1.5 wt% and (e) 2.0 wt%. | |
4. Conclusions
A novel 2-D transition metal carbide MXene (nanolayered Ti2CTx) was synthesised by etching a MAX phase in the liquid phase and subsequently structurally homogeneous Ti2CTx/EP nanocomposites were prepared by in situ intercalative polymerization. As shown by XRD and SEM, EP was intercalated into the Ti2CTx interlayer and the exfoliated Ti2CTx nanosheets were dispersed evenly in the EP matrix. The results of DMA and TGA revealed that the incorporation of Ti2CTx nanosheets into the EP matrix significantly increased the Tg and Es values and improved the creep resistance and recovery properties, but reduced the thermal stability of EP. Compared with those of pure EP, the Tg and Es values increased by 15% and 38%, respectively. In addition, an increase in the applied level of creep stress increased the level of creep deformation of the composites, which displayed much lower levels of creep strain with an increase in the Ti2CTx content. Furthermore, the formation of an interface and a crosslinked network structure had an advantageous effect on the mechanical properties. Although the surface microhardness of the composites decreased slightly, the impact strength and flexural strength increased by 76% and 66%, respectively, when the content of added Ti2CTx was 1.0 wt%. In particular, tribological studies revealed a lower friction coefficient at a high mass concentration of Ti2CTx and a higher sliding velocity than that of pure EP. The worn surface of the Ti2CTx/EP composite was smoother than that of pure EP, which suggested that severe wear damage was mitigated and the Ti2CTx nanosheets had a positive effect on solid lubrication.
Acknowledgements
The authors gratefully acknowledge the support of the National Natural Science Foundation of China (No. 51205111, 51472075 and 51275156), Plan for Scientific Innovation Talent of Henan Province (No. 134100510008), Program for Innovative Research Team of Henan Polytechnic University (T2013-4), and Opening Project of Henan Key Discipline Open Laboratory of Mining Engineering Materials (MEM12-11).
References
- L. X. Gong, L. Zhao, L. C. Tang, H. Y. Liu and Y. W. Mai, Compos. Sci. Technol., 2015, 121, 104–114 CrossRef CAS
. - S. S. Vaisakh, K. V. Mahesh, S. Balanand, R. Metz, M. Hassanzadeh and S. Ananthakumar, RSC Adv., 2015, 5, 16521–16531 RSC
. - A. A. Azeez, K. Y. Rhee, S. J. Park and D. Hui, Composites, Part B, 2013, 45, 308–320 CrossRef CAS
. - B. Mayoral, E. Harkin Jones, P. Noorunnisa Khanam, M. A. AlMaadeed, M. Ouederni, A. Hamiltona and D. Sun, RSC Adv., 2015, 5, 52395–52409 RSC
. - K. R. Reddy, K. P. Lee, A. I. Gopalan and H. D. Kang, React. Funct. Polym., 2007, 67, 943–954 CrossRef CAS
. - K. R. Reddy, K. P. Lee, A. I. Gopalan and S. A. Md, Polym. Adv. Technol., 2007, 18, 38–43 CrossRef CAS
. - S. H. Choi, D. H. Kim, A. V. Raghu, K. R. Reddy, H.-I. Lee, K. S. Yoon, H. M. Jeong and B. K. Kim, J. Macromol. Sci., Part B: Phys., 2012, 51, 197–207 CrossRef CAS
. - Y. P. Zhang, S. H. Lee, K. R. Reddy, A. I. Gopalan and K. P. Lee, J. Appl. Polym. Sci., 2007, 104, 2743–2750 CrossRef CAS
. - K. R. Reddy, B. C. Sin, C. H. Yoo, D. Sohn and Y. Lee, J. Colloid Interface Sci., 2009, 340, 160–165 CrossRef CAS PubMed
. - K. R. Reddy, B. C. Sin, K. S. Ryu, J. C. Kim, H. Chung and Y. Lee, Synth. Met., 2009, 159, 595–603 CrossRef CAS
. - H. Y. Liu, G. T. Wang, Y. W. Mai and Y. Zeng, Composites, Part B, 2011, 42, 2170–2175 CrossRef
. - M. Martin-Gallego, R. Verdejo, M. A. Lopez-Manchado and M. Sangermano, Polymer, 2011, 52, 4664–4669 CrossRef CAS
. - J. Kim, B. s. Yim, J. m. Kim and J. Kim, Microelectron. Reliab., 2012, 52, 595–602 CrossRef CAS
. - L. C. Tang, Y. J. Wan, D. Yan, Y. B. Pei, L. Zhao, Y. B. Li, L. B. Wu, J. X. Jiang and G. Q. Lai, Carbon, 2013, 60, 16–27 CrossRef CAS
. - I. Zaman, Q. H. Le, H. C. Kuan, N. Kawashima, L. Luong, A. Gerson and J. Ma, Polymer, 2011, 52, 497–504 CrossRef CAS
. - X. Li, Y. Gao, J. Xing, Y. Wang and L. Fang, Wear, 2004, 257, 279–283 CrossRef CAS
. - O. Eksik, J. Gao, S. A. Shojaee, A. Thomas, P. Chow, S. F. Bartolucci, D. A. Lucca and N. Koratkar, ACS Nano, 2014, 8, 5282–5289 CrossRef CAS PubMed
. - J. Yu, X. Huang, C. Wu, X. Wu, G. Wang and P. Jiang, Polymer, 2012, 53, 471–480 CrossRef CAS
. - S. Sinha Ray and M. Okamoto, Prog. Polym. Sci., 2003, 28, 1539–1641 CrossRef
. - Q. H. Zeng, A. B. Yu, G. Q. Lu and D. R. Paul, J. Nanosci. Nanotechnol., 2005, 5, 1574–1592 CrossRef CAS PubMed
. - S. G. Prolongo, A. Jimenez-Suarez, R. Moriche and A. Ureña, Compos. Sci. Technol., 2013, 86, 185–191 CrossRef CAS
. - P. Meneghetti and S. Qutubuddin, J. Colloid Interface Sci., 2005, 288, 387–389 CrossRef CAS PubMed
. - N. Michael, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi and M. W. Barsoum, Adv. Mater., 2011, 23, 4248–4253 CrossRef PubMed
. - I. R. Shein and A. L. Ivanovskii, Comput. Mater. Sci., 2012, 65, 104–114 CrossRef CAS
. - K. J. Harris, M. Bugnet, M. Naguib, M. W. Barsoum and G. R. Goward, J. Phys. Chem. C, 2015, 119, 13713–13720 CAS
. - M. W. Barsoum, Prog. Solid State Chem., 2000, 28, 201–281 CrossRef CAS
. - J. C. Lei, X. Zhang and Z. Zhou, Front. Phys. China, 2015, 10, 276–286 Search PubMed
. - O. Mashtalir, K. M. Cook, V. N. Mochalin, M. Crowe, M. W. Barsoum and Y. Gogotsi, J. Mater. Chem. A, 2014, 2, 14334–14338 CAS
. - D. Sun, M. Wang, Z. Li, G. Fan, L.-Z. Fan and A. Zhou, Electrochem. Commun., 2014, 47, 80–83 CrossRef CAS
. - M. Naguib, V. N. Mochalin, M. W. Barsoum and Y. Gogotsi, Adv. Mater., 2014, 26, 992–1005 CrossRef CAS PubMed
. - Y. Gao, L. Wang, Z. Li, A. Zhou, Q. Hu and X. Cao, Solid State Sci., 2014, 35, 62–65 CrossRef CAS
. - M. R. Lukatskaya, O. Mashtalir, C. E. Ren, Y. Dall'Agnese, P. Rozier, P. L. Taberna, M. Naguib, P. Simon, M. W. Barsoum and Y. Gogotsi, Science, 2013, 341, 1502–1505 CrossRef CAS PubMed
. - A. N. Enyashin and A. L. Ivanoyskii, J. Phys. Chem. C, 2013, 117, 13637–13643 CAS
. - A. N. Enyashin and A. L. Ivanovskii, J. Theor. Comput. Chem., 2012, 989, 27–32 CrossRef CAS
. - S. Zhao, W. Kang and J. Xue, Appl. Phys. Lett., 2014, 104, 133106 CrossRef
. - Q. Peng, J. Guo, Q. Zhang, J. Xiang, B. Liu, A. Zhou, R. Liu and Y. Tian, J. Am. Chem. Soc., 2014, 136, 4113–4116 CrossRef CAS PubMed
. - S. Zhao, W. Kang and J. Xue, J. Phys. Chem. C, 2014, 118, 14983–14990 CAS
. - X. Liang, A. Garsuch and L. F. Nazar, Angew. Chem., Int. Ed., 2015, 54, 3907–3911 CrossRef CAS PubMed
. - J. Come, M. Naguib, P. Rozier, M. W. Barsoum, Y. Gogotsi, P. L. Taberna, M. Morcrette and P. Simon, J. Electrochem. Soc., 2012, 159, A1368–A1373 CrossRef CAS
. - D. Er, J. Li, M. Naguib, Y. Gogotsi and V. B. Shenoy, ACS Appl. Mater. Interfaces, 2014, 6, 11173–11179 CAS
. - M. Ghidiu, M. R. Lukatskaya, M.-Q. Zhao, Y. Gogotsi and M. W. Barsoum, Nature, 2014, 516, 78–U171 CAS
. - X. Zhang, M. Xue, X. Yang, Z. Wang, G. Luo, Z. Huang, X. Sui and C. Li, RSC Adv., 2015, 5, 2762–2767 RSC
. - J. Yang, B. Chen, H. Song, H. Tang and C. Li, Cryst. Res. Technol., 2014, 49, 926–932 CrossRef CAS
. - S. Gupta, T. Hammann, R. Johnson and M. F. Riyad, Tribol. Trans., 2015, 58, 560–566 CrossRef CAS
. - H. Zhang, L. Wang, Q. Chen, P. Li, A. Zhou, X. Cao and Q. Hu, Mater. Des., 2016, 92, 682–689 CrossRef CAS
. - S. Y. Yang, W. N. Lin, Y. L. Huang, H. W. Tien, J. Y. Wang, C. C. M. Ma, S. M. Li and Y. S. Wang, Carbon, 2011, 49, 793–803 CrossRef CAS
. - L. Li, A. Zhou, L. Xu, Z. Li and L. Wang, J. Wuhan Univ. Technol., Mater. Sci. Ed., 2013, 28, 882–887 CrossRef CAS
. - M. Ghidiu, M. R. Lukatskaya, M. Q. Zhao, Y. Gogotsi and M. W. Barsoum, Nature, 2014, 516, 78–81 CAS
. - M. Naguib, M. Olha, J. Carle, V. Presser, J. Lu, L. Hultman, Y. Gogotsi and M. W. Barsoum, ACS Nano, 2012, 6, 1322–1331 CrossRef CAS PubMed
. - Y. R. Lee, S. C. Kim, H.-i. Lee, H. M. Jeong, A. V. Raghu, K. R. Reddy and B. K. Kim, Macromol. Res., 2011, 19, 66–71 CrossRef CAS
. - M. Wang, X. Fan, W. Thitsartarn and C. He, Polymer, 2015, 58, 43–52 CrossRef CAS
. - L. C. Tang, H. Zhang, S. Sprenger, L. Ye and Z. Zhang, Compos. Sci. Technol., 2012, 72, 558–565 CrossRef CAS
. - S. Chandrasekaran, N. Sato, F. Tölle, R. Mülhaupt, B. Fiedler and K. Schulte, Compos. Sci. Technol., 2014, 97, 90–99 CrossRef CAS
. - M. A. Rafiee, J. Rafiee, I. Srivastava, Z. Wang, H. Song, Z.-Z. Yu and N. Koratkar, Small, 2010, 6, 179–183 CrossRef CAS PubMed
. - F. Gojny, M. Wichmann, B. Fiedler and K. Schulte, Compos. Sci. Technol., 2005, 65, 2300–2313 CrossRef CAS
. - J. Ma, M. S. Mo, X. S. Du, P. Rosso, K. Friedrich and H. C. Kuan, Polymer, 2008, 49, 3510–3523 CrossRef CAS
. - L. Le Pluart, J. Duchet and H. Sautereau, Polymer, 2005, 46, 12267–12278 CrossRef CAS
. - I. Zaman, T. T. Phan, H.-C. Kuan, Q. Meng, L. T. Bao La, L. Luong, O. Youssf and J. Ma, Polymer, 2011, 52, 1603–1611 CrossRef CAS
. - J. Li, Y. Du, C. Huo, S. Wang and C. Cui, Ceram. Int., 2015, 41, 2631–2635 CrossRef CAS
. - S. Jin Han, H.-I. Lee, H. M. Jeong, B. K. Kim, A. V. Raghu and K. R. Reddy, J. Macromol. Sci., Part B: Phys., 2014, 53, 1193–1204 CrossRef
. - L. C. Tang, X. Wang, L. X. Gong, K. Peng, L. Zhao, Q. Chen, L. B. Wu, J. X. Jiang and G. Q. Lai, Compos. Sci. Technol., 2014, 91, 63–70 CrossRef CAS
. - A. Zandiatashbar, C. R. Picu and N. Koratkar, Small, 2012, 8, 1676–1682 CrossRef CAS PubMed
. - J. Shen, W. Huang, L. Wu, Y. Hu and M. Ye, Compos. Sci. Technol., 2007, 67, 3041–3050 CrossRef CAS
. - L. Tang, H. Zhang, X. Wu and Z. Zhang, Polymer, 2011, 52, 2070–2074 CrossRef CAS
. - J. Yang, Z. Zhang, K. Friedrich and A. K. Schlarb, Macromol. Rapid Commun., 2007, 28, 955–961 CrossRef CAS
. - T. Ramanathan, A. A. Abdala, S. Stankovich, D. A. Dikin, M. Herrera-Alonso, R. D. Piner, D. H. Adamson, H. C. Schniepp, X. D. Chen, R. S. Rruoff, S. T. Nguyen, I. A. Aksay, R. K. Prud’Homme and D. L. C. Brinson, Nat. Nanotechnol., 2008, 3, 327–331 CrossRef CAS PubMed
. - L. G. X. Wang, L. C. Tang, K. Peng, Y. B. Pei, L. Zhao, L. B. Wu and J. X. Jiang, Composites, Part A, 2015, 69, 288–298 CrossRef
. - W. Li, A. Dichiara and J. Bai, Compos. Sci. Technol., 2013, 74, 221–227 CrossRef CAS
. - X. Wang, W. Xing, P. Zhang, L. Song, H. Yang and Y. Hu, Compos. Sci. Technol., 2012, 72, 737–743 CrossRef CAS
. - W. J. Wang, C. W. Li and K. P. Chen, Mater. Lett., 2013, 110, 61–64 CrossRef CAS
. - R. K. Prusty, D. K. Rathore, M. J. Shukla and B. C. Ray, Composites, Part B, 2015, 83, 166–174 CrossRef CAS
. - H. Jin, G. M. Miller, N. R. Sottos and S. R. White, Polymer, 2011, 52, 1628–1634 CrossRef CAS
. - S. Laurenzi, S. Botti, A. Rufoloni and M. G. Santonicola, Procedia Eng., 2014, 88, 157–164 CrossRef CAS
. - A. Dasari, Z. Z. Yu and Y. W. Mai, Mater. Sci. Eng., R, 2009, 63, 31–80 CrossRef
. - Y. Luo, M. Z. Rong and M. Q. Zhang, Polym. Polym. Compos., 2005, 13, 245–252 CAS
.
|
| This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.