Vous êtes sur la page 1sur 6

A.

Zandiatashbar
Graduate Research Assistant Student Mem. ASME

R. C. Picu1
Professor Fellow ASME e-mail: picuc@rpi.edu

Mechanical Behavior of Epoxy-Graphene Platelets Nanocomposites


Various aspects of the mechanical behavior of epoxy-based nanocomposites with graphene platelets (GPL) as additives are discussed in this article. The monotonic loading response indicates that at elevated temperatures, the elastic modulus and the yield stress are signicantly improved in the composite as compared to neat epoxy. The activation energy for creep is smaller in neat epoxy, which indicates that the composite creeps less, especially at elevated temperatures and higher stresses. The composites also exhibit larger fracture toughness. When subjected to cyclic loading, fatigue crack growth rate is smaller in the composite relative to neat epoxy. This reduction is important by at least an order of magnitude at all stress intensity factor amplitudes. Optimal property improvements in the monotonic, cyclic, and fracture behaviors are obtained for very low lling fraction of approximately 0.1 wt. %. Similar differences in the mechanical behavior are observed when the composite is probed on the local scale by nanoindentation. [DOI: 10.1115/1.4006499]

N. Koratkar
Professor Department of Mechanical, Aerospace and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180

Introduction

Polymer nanocomposites received signicant attention over the last decade due to multiple experimental observations that signicant enhancement of mechanical properties can be obtained by the addition of small volume fractions of nanoscale additives [1,2]. It was seen that the polymers lled with only few percent volume fraction of functionalized nanoparticles have signicantly larger strength and ductility than the corresponding neat material [3]. When carbon nanotube (CNT) additives are used, enhanced properties are obtained at even smaller lling fractions, below 1 wt. % [47]. Graphene has emerged recently as a promising new nanoscale ller material [8,9]. However, due to the cost of producing large quantities of monolayer graphene as well as issues related to their stability in powder form and during processing, the nanocomposites research community has employed primarily GPL [1]. These are thin plates, each containing several graphene sheets. The platelets have in-plane dimensions about 2 orders of magnitude larger than the thickness, large in-plane stiffness and relatively a low bending stiffness. In many cases, these contain a large concentration of defects which reduce, to some extent, the stiffness and strength of llers, but facilitate functionalization and bonding to the polymeric matrix. The GPL has been used to reinforce various polymers, both cross-linked, such as epoxy [10], and uncross-linked, such as PMMA [11]. The effect of lling varies signicantly as a function of the lling fraction, the functionalization of the ller surface, dispersion and distribution of llers in the matrix. Generally, a necessary (but insufcient) condition for obtaining enhanced composite properties is to reach proper ller dispersion. If this condition is achieved, a very small ller volume fraction may be sufcient. The nature of the ller-matrix interface is also of importance; weak interfaces promote cavitation, which in turn enhances the material ductility. As in regular composites with microscale llers, strong ller-matrix adhesion is required for strength improvements. Fracture toughness is enhanced for intermediate values of the interface strength.
1 Corresponding author. Contributed by the Materials Division of ASME for publication in the JOURNAL OF ENGINEERING MATERIALS AND TECHNOLOGY. Manuscript received October 16, 2011; nal manuscript received January 7, 2012; published online May 16, 2012. Assoc. Editor: Xin-Lin Gao.

This work addresses various aspects of the mechanical behavior of epoxy lled with GPL. Epoxy is an important technological material with applications ranging from adhesives used in civil engineering and the microelectronic industry, to matrix material for composite helicopter blades. In all these applications strength and toughness, and in particular fatigue resistance, are crucial. We have studied both the monotonic and cyclic loading response of epoxy-GPL composites, at room and elevated temperatures (below the glass transition temperature) [12]. Creep deformation was also studied, as it is important, for example, in civil engineering applications of epoxy. We have shown that when llers are properly dispersed in the matrix, 0.1 wt. % GPL is sufcient to lead to optimal mechanical properties. In this article, we review fatigue, creep, and fracture toughness results obtained for this system. New results on the monotonic and nanoindentation behavior of epoxy-GPL are presented, completing the characterization of this class of materials. The article is organized as follows: the materials and methods used for preparation and characterization are presented in Sec. 2; Sec. 3 is divided in subsections addressing the characterization of the ller dispersion in the matrix, monotonic test results, creep, fracture, and fatigue behavior, as well as a comparison with the nanoscale material response obtained by nanoindentation. The conclusions are presented in Sec. 4.

Materials and Methods

Bulk quantities of graphene platelets were prepared by rapid thermal exfoliation of graphite oxide at a heating rate greater than 2000  C/min [12]. Graphite oxide was obtained by oxidizing graphite akes in a solution of nitric acid, sulfuric acid, and potassium chlorate for about 96 h [9,13]. For thermal exfoliation, graphite oxide was placed in a quartz tube, which was sealed at one end and closed by using a rubber stopper at the other end. An argon inlet was embedded through the stopper, and the sample was ushed by argon for 10 min. Then, the tube is inserted in a tube furnace (79300, Thermo Fisher Scientic, Inc., USA) preheated to 1050  C and held there for about 30 s. Graphene obtained by this method was in the form of GPL [12]. The GPLs were dispersed in acetone (100 ml of acetone to 0.1 g of GPL) using a sonicator at high amplitude (Sonics Vibracell VC 750, Sonics and Materials, Inc., USA), for 1.5 h. The epoxy resin (System 2000 Epoxy Resin, Fibreglast, Inc., USA) was JULY 2012, Vol. 134 / 031011-1

Journal of Engineering Materials and Technology C 2012 by ASME Copyright V

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

added to the mixture and sonicated following the same procedure for another 1.5 h. Next, the acetone was evaporated by heating the mixture on a magnetic stir plate using a Teon-coated magnetic bar for 3 h at 70  C. The mixture was placed in a vacuum chamber for 12 h at 70  C to ensure that all the acetone has been removed. After allowing the GPL/epoxy resin slurry to cool down to room temperature, a low viscosity curing agent (2120 Epoxy Hardener, Fibreglast, Inc., USA) was added and mixed using a high speed shear mixer (ARE-250, Thinky, Japan) for 4 min at 2000 rpm. The mixture was again placed in a vacuum chamber for degassing for approximately 30 min. Finally, the mixture was poured into silicone molds and was cured at room temperature and 90 psi pressure for 24 h, followed by 4 h of postcure at 90  C. The samples were used for mechanical testing which was performed with a standard material testing machine (MTS-858, MTS Systems Corp., MN, USA) tted with an environmental chamber. Monotonic, creep, and fatigue tests were performed with the same machine. To evaluate the composites performance under fatigue conditions, dynamic crack propagation tests on compact tension samples were conducted following ASTM E647 standard. Nanoindentation was performed using a NanoTestV 550 (Micro Materials Ltd., Wrexham, UK) on nanocomposite thin lms. The lm material was obtained by a procedure similar to that described above, except that the mixture was spun on a Si wafer before curing. The curing process was similar to that for bulk specimens.
R

0.5 wt. % GPL. The distribution for the composite with larger lling fraction is shifted to larger values of the variable, indicating that a larger degree of GPL clustering takes place. Some GPL clusters are also present in the 0.1 wt. % sample, but the mean of that distribution is at 1 lm, i.e., most inclusions are submicron in size and likely few-layer graphene platelets. This suggests that better mixing is obtained at lower lling fractions, and that large clusters of GPL are present for lling fractions larger than approximately 0.5 wt. %. Since interactions within a cluster are rather weak (van der Waals), these inclusions have low strength and modulus and are expected to lead to worse mechanical properties of the composite relative to materials with low lling and even relative to neat epoxy. This effect is conrmed experimentally, as discussed below. 3.2 Uniaxial Monotonic Testing. Monotonic uniaxial tensile testing was performed using dog-bone samples. Figure 2(a) shows stressstrain curves for neat epoxy and two composites, with 0.1 and 0.2 wt. %. Tests were performed at room temperature and at 40  C, 55  C, and 70  C. At room temperature, the curves corresponding to the three materials are close to each other. The stiffness increases slightly, mostly proportional to the volume fraction of inclusions. Note that the relationship between weight and volume fraction in this system is 1 wt. % 1.12 vol. % [10]. The yield stress is slightly larger in composites relative to the neat epoxy, but the difference is marginally larger than the variability expected from sample to sample and experimental uncertainties. The curves reach a maximum at a strain of about 4%. Note that previous studies [10,12] have indicated 2030% to increase in modulus and strength of epoxies at GPL loading fractions of ~0.1 wt. % at room temperature. These differences may arise due to the variation in the properties of the GPL used (e.g., defect density, thickness, etc.). Signicant changes in the stressstrain curve are observed as the temperature increases. As expected, the 0.2% yield stress decreases and ductility (strain at failure) increases. However, the decrease is more pronounced in neat epoxy. Figure 2(b) shows the magnitude of the yield stress of the composite relative to that of neat epoxy as a function of temperature. The relative magnitude is repoxy =repoxy and is given in Fig. 2(b) in evaluated as rcomp y y y percentage. Note the signicant increase of this measure as the

Results and Discussion

3.1 Nanofiller Size Distribution and Dispersion. Nanocomposite properties depend strongly on the quality of dispersion of llers in the polymeric matrix. The dispersion is achieved during the mixing process; it is considered that the cross-linking process introduces no bias in this measure. Figure 1(a) shows an image of the fracture surface of a composite with 0.1 wt. % GPL. The image shows that the sample preparation procedure described in Sec. 2 leads to good ller dispersion. However, not all llers are of the same size, since GPL clustering in the powder before mixing is pronounced. The probability distribution function of inclusion sizes was determined by processing images of fracture surfaces and using the line intercept method (ASTM E112). The distribution is shown in Fig. 1(b) for two composites with 0.1 and

Fig. 1 (a) SEM micrograph of a fracture surface of the 0.1 wt. % epoxy-GPL nanocomposite after failure. The inset shows a region of the main image at higher magnication. The inclusions can be easily distinguished from the river pattern features, as indicated in the inset. (b) Probability distribution function of inclusion size in 0.1 and 0.5 wt. % epoxy-GPL nanocomposites obtained by image processing of fracture surface micrographs [14].

031011-2 / Vol. 134, JULY 2012

Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

Fig. 2 (a) Engineering stress versus engineering strain plots for neat epoxy, 0.1 and 0.2 wt. % epoxy-GPL nanocomposites. Symbols are added to the curve just for labeling. (b) Relative magnitude of the 0.2% yield stresses of the composite and neat repoxy =repoxy ), versus temperature. Data for composites with 0.1 and 0.2 wt. % GPL are shown. epoxy (computed as rcomp y y y The nonmonotonic variation of the GPL0.1 curve is expected to be due to the variability from sample to sample. However, the increasing trend of the measure shown is clear and independent of the measurement noise.

temperature increases. It is also noted that the 0.1 wt. % composite provides better performance than the composite with 0.2 wt. % GPL. This is also observed and discussed in Sec. 3.4 in connection with fracture toughness. Increasing temperature also leads to a dramatic decrease of the apparent Youngs modulus, with the drop being much more pronounced in neat epoxy. Although the modulus is expected to decrease with increasing temperature, the signicant variation observed here is likely an effect of the pronounced transient creep at elevated temperatures, as discussed in Sec. 3.3. An interesting observation is that as the temperature increases, the strain hardening increases in all materials. This is against expectations based on the behavior of most engineering materials. A strong strain hardening promotes stable plastic ow, enhanced ductility and toughness. Also interestingly, the strain hardening rate is almost a constant of strain at all temperatures. Very little (if any) necking is observed in these materials at all temperatures considered. The three materials have the same strain hardening rate at given temperature. It is appropriate to discuss these observations while taking into account the difference between the test temperature and the glass transition temperature, Tg. Tg was measured using differential scanning calorimetry for neat epoxy and nanocomposites of 0.1, 0.3, and 0.5 wt. % GPL. The values range between 73 and 79  C for all these materials. While the mean values of Tg seem to increase with the ller content, the experimental noise is rather large (62  C) and a clear correlation cannot be established. However, the observations in Fig. 2, i.e., the faster softening of neat epoxy, cannot be explained based on the small difference of Tg between the various materials, primarily because this argument predicts the opposite trend relative to the experiment. At given temperature, the nanocomposite is closer to its Tg than unlled epoxy. Note that the degradation temperature of epoxy is above 170  C. The strain rate sensitivity of the three materials was measured by performing two monotonic tension tests at strain rates 105 s1 and 104 s1, i.e., with a strain rate differential of 10, and at room temperature. The strain rate sensitivity is positive and increases with strain. At the peak stress (4% strain), the strain rate sensitivity parameter is m 0.057. The parameter is evaluated as _1 =e _2 , with the two values of the stress, m logr1 =r2 =loge r1 and r2, being measured at same strain and at the two strain _2 . _1 and e rates, e 3.3 Creep. Creep tests were performed with nominal stresses of 20 and 40 MPa at room temperature, in separate experiments Journal of Engineering Materials and Technology

[14]. These stress levels represent 35% and 70% of the yield stress, respectively. The load was increased fast and then held constant. The creep strain was recorded for 36 h. No necking or signicant variation of the specimen cross-sectional area was observed during the test, so the effective stress is that prescribed throughout the entire test. Figure 3 shows the creep strain (total strain minus the elastic strain corresponding to the creep load) versus time for neat epoxy and for the composite with 0.1 wt. % GPL at room temperature and nominal stress of 40 MPa. Tests were performed with composites having higher lling fractions, up to 0.5 wt. %, but the 0.1 wt. % composite exhibited the lowest creep rate. In fact, the 0.5 wt. % GPL composite exhibits larger creep strains than neat epoxy at all times. After the transient creep period (which lasts for approximately 1 h at this stress), steady state creep sets in and the curve is approximated by e(t) $ t1/q, Fig. 3. The exponent q is 2.49 for epoxy and 4.71 for the nanocomposite, which indicates that, as time increases, a smaller creep strain is recorded for the composite. The power law dependence of the creep strain on time can be captured by a fractional viscoelastic model, a model introduced to represent long tails, power law relaxation of complex liquids and solids [15]. A power function of this type is usually interpreted as

Fig. 3 Loglog plot of creep strain versus time (in seconds) for neat epoxy and 0.1 wt. % epoxy-GPL showing a power law creep compliance function [14]

JULY 2012, Vol. 134 / 031011-3

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

Fig. 4 Creep strain versus time for neat epoxy and 0.1 wt. % epoxy-GPL at a stress of 20 MPa and temperatures of 23, and 55  C [14]

Fig. 5 Fracture toughness values for neat epoxy, 0.1, and 0.2 wt. % epoxy-GPL nanocomposites at room temperature, 55, and 70  C

a manifestation of hierarchical relaxation processes taking place in the material. Figure 4 shows the effect of temperature on creep of neat epoxy and 0.1 wt. % GPL composite at 20 MPa nominal stress. As the temperature increases, the difference between the creep strains of neat epoxy and the composite increases, becoming quite important (~50% difference) at 55  C. The composite creeps less in all cases. This behavior is in agreement with the monotonic response shown in Fig. 2, where it has been seen that neat epoxy loses its carrying capacity much faster than the GPL composites as the temperature increases above ambient conditions. The result shown in Fig. 4 is important as the reduction of the creep rate happens in conditions in which creep resistance is more critical, i.e., at high temperatures. 3.4 Fracture and Fatigue Behavior. The fracture behavior as well as the resistance to fatigue of these materials has been studied and discussed in several publications [10,12]. Here, we present a brief review of the main results and provide new results regarding the variation of toughness with the temperature. A discussion comparing epoxy-GPL and epoxy-CNT nanocomposites is included in order to put in perspective, the behavior of epoxyGPL on which this article is focused. The fracture toughness of composites with various fractions of GPL was determined at several temperatures, following the ASTM D5045 standard. Figure 5 shows the average fracture toughness values for each material at room and elevated temperatures. Three samples have been tested for each condition. As shown in previous reports [12], 0.1 wt. % is the optimum ller content at which the toughness values are increased the most. By increasing the temperature, the toughness values increase for all materials, while the difference between epoxy and the two composites considered here becomes smaller. Insight into this behavior can be obtained by comparing the areas under the curves in Fig. 2(a). As the temperature increases, the ow stress decreases, but the failure strain increases. These effects compensate each other leading to the observed apparent temperature insensitivity of the toughness seen in Fig. 5. The results of fatigue tests are shown in Fig. 6 [10] for neat epoxy and for epoxy-GPL composite with 0.1 wt. % ller content. Data for epoxy-single wall and multiwall carbon nanotubes of same lling fraction are reproduced from Ref. [10], for reference. The gure shows the crack growth rate, da/dN, versus the stress intensity factor amplitude and demonstrates that nanollers increase the fatigue resistance dramatically. The crack growth rate is an order of magnitude smaller in the epoxy-GPL composite relative to neat epoxy. A similarly important reduction is observed in 031011-4 / Vol. 134, JULY 2012

Fig. 6 Fatigue crack growth rate (da/dN) versus stress intensity factor amplitude (DK) for neat epoxy and 0.1 wt. % lled epoxy with SWNT, MWNT, and GPL [10]

epoxy-CNT composites, but the slowdown is lost when the stress intensity factor amplitude increases. This is not observed in epoxy-GPL, which makes this composite more useful for fatigue applications than its counterpart containing CNTs. It is also observed that no signicant difference exists between single wall and multiwall CNT composites. The toughening mechanism in epoxy-CNT composites was identied as crack bridging. This mechanism leads to toughening in regular ber composites and was observed in Ref. [16] to also operate at the nanoscale. CNTs bridge the crack faces and are being pulled-out, therefore providing an additional dissipation mechanism. As the stress intensity factor increases, the crack opening displacement increases and the size of the region in the vicinity of the crack tip in which pull-out processes take place decreases. This is captured by the model developed in Ref. [17] for regular ber composites, and used in Ref. [16] for this nanocomposite system. The mechanism of toughening in epoxy-GPL is not fully understood at present; however, it is expected that GPL modies the plastic ow and dissipation in the crack plastic zone, while some form of crack pinning and deection should also be active. Some experimental proof for the latest mechanism in the form of a correlation between the roughness of the fracture surface and the fracture toughness has been discussed in Ref. [12], but a mechanistic understanding of the process is still lacking. Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

Fig. 7 (a) Top-view optical image of the surface of a 0.1 wt. % epoxy-GPL nanocomposite lm. Sites where nanoindentation is performed are indicated. (b) AFM 3D image of GPL protrusions on the lm surface after being etched with acetone.

The effect of the CNT length and diameter on the crack growth rate under fatigue loading has been discussed in Ref. [18]. It was shown that the crack growth rate decreases as the CNT diameter decreases and as the length increases. These conclusions are in line with the crack bridging model used in Ref. [16]. 3.5 Probing Local Mechanical Behavior by Nanoindentation. Nanoscale mechanical properties are investigated by nanoindentation. Tests are performed using 0.1 wt. % epoxyGPL composite lms prepared as described in Sec. 2. The lms are spun on a Si substrate using a Headway spinner at 8000 rpm spinning speed for 40 s. The lms are approximately 80 lm thick. Due to the small thickness, they are transparent and the dispersion and distribution of llers can be observed by optical microscopy. Figure 7(a) shows a top view of the lm, with many clusters of dimensions ranging from 1 to 5 lm being visible. A large number of smaller, submicron inclusions, likely few-layer GPLs, are visible at higher magnication and by AFM probing of the surface. Most inclusions are below the surface. AFM topography indicates that the surface roughness is very small, approximately 10 nm. To gain access to inclusions, the lm was etched by immersion in acetone followed by immersion in isopropylene. Samples were then dried by blowing nitrogen. This process exposes inclusions and clusters, as shown in Fig. 7(b).

The molecular scale structure of these lms is likely similar to that of the bulk samples since (a) the lms are much thicker than the chain gyration radius (on the order of nanometer), and (b) the lms were annealed sufciently during curing such that the molecular structure is expected to be close to equilibrium. Furthermore, images of the ller distribution do not reveal any preferential ller alignment which may have been induced by spinning. Using optical microscopy and AFM topography information, one may select locations for indentation. Specically, sites away from inclusions and sites on protruding llers were selected in separate experiments. Indentation was performed using a NanoTestV nanoindenter from Micro Materials Ltd. and a Berkovich diamond indenter. Figure 8(a) shows multiple indentation curves, including loadingunloading and cyclic loading with incomplete unloading and reloading up to increasing levels of load. In all cases, a hold period of 10 s is applied at the maximum load. These curves were obtained by indenting at the sites shown in Fig. 7(a) and labeled as single-cycle (i.e., simple loadingunloading up to 8 mN load), and multiple-cycle (cyclic loadingunloading with the maximum load of 8 mN). These curves are shown with continuous lines in Fig. 8(a). All sites were away from protruding inclusions or from subsurface inclusions which were visible optically.
R

Fig. 8 (a) Load-displacement curves for single-cycle and multiple-cycle nanoindentation in the neat epoxy lm (solid line) and the nanocomposite lm (dashed line). The plot shows a good repeatability from indentation to indentation. (b) Multiple-cycle indentation in the nanocomposite lm, at and away from GPL protrusions (Fig. 7(b)) [20].

Journal of Engineering Materials and Technology

JULY 2012, Vol. 134 / 031011-5

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

Hence, the response corresponds to the epoxy matrix. It is seen that the repeatability (site to site) is very good. The cyclic test is load controlled with displacement limits, and the loading branches are seen to follow closely the monotonic curves. Some hysteresis, indicating some degree of viscoplasticity is seen. Figure 8(a) includes cyclic curves obtained with neat epoxy samples prepared under the same conditions. As in the case of the composite, the repeatability from site to site is an excellent, and the hysteresis loops and dissipation are similar to those observed in the composite. The neat epoxy curves overlap with those of the composite at small loads, which indicates that we are probing locally epoxy which is not modied by the presence of inclusions. As the load and indentation depth increase, the composite appears stiffer. This indicates that, as larger loads are applied and a larger plastic zone develops around the indenter, llers play a role and increase the ow stress (see also Fig. 2). Furthermore, the present results do not provide support for the existence of a bonded polymer layer in these composites. Such a layer with modied mechanical and dielectric properties was postulated to exist in all nanolled thermoplastics. The modied macroscopic properties of the nanocomposite were associated with the percolation of these modied layers. In the materials studied here, the interphase is thought to be less well dened due to the presence of cross-links [19]. Figure 8(b) shows cyclic indentation curves obtained by probing on protruding llers and away from those, in the composite. As expected, the response to indentations on protrusions is much stiffer than that to indentations elsewhere in the epoxy matrix (these curves are identical to those in Fig. 8(a)). The local mechanical tests provide results which are in agreement with the trends seen macroscopically. Although they do not clarify the mechanism of toughening observed in monotonic and cyclic tests, they indicate that (a) GPLs do not have a long-range effect on the epoxy matrix, as postulated in nanolled thermoplastics, and most of the epoxy in the composite has the same behavior as unlled epoxy, (b) GPLs have a strengthening effect, presumably due to the good bonding with the epoxy matrix (mediated by the O and OH groups present on the surface of GPL) [20].

Acknowledgment
We gratefully acknowledge support from the NSF under Grant No. CMMI 0900188 and from the ONR under Grant No. N0001409-1-0928.

References
[1] Potts, J. R., Dreyer, D. R., Bielawski, C. W., and Ruoff, R. S., 2011, Graphene-Based Polymer Nanocomposites, Polymer, 52(1), pp. 525. [2] Li, B., and Zhong, W.-H., 2011, Review on Polymer/Graphite Nanoplatelet Nanocomposites, J. Mater. Sci., 46(17), pp. 55955614. [3] Ash, B. J., Siegel, R. W., and Schadler, L. S., 2004, Mechanical Behavior of Alumina/Poly(Methyl Methacrylate) Nanocomposites, Macromolecules, 37(4), pp. 13581369. [4] Coleman, J., Khan, U., Blau, W., and Gunko, Y., 2006, Small But Strong: A Review of the Mechanical Properties of Carbon Nanotubepolymer Composites, Carbon, 44(9), pp. 16241652. [5] Ajayan, P. M., Schadler, L. S., Giannaris, C., and Rubio, A., 2000, SingleWalled Carbon Nanotube-Polymer Composites: Strength and Weakness, Adv Mater, 12(10), pp. 750753. [6] Schadler, L. S., Giannaris, S. C., and Ajayan, P. M., 1998, Load Transfer in Carbon Nanotube Epoxy Composites, Appl. Phys. Lett., 73(26), p. 3842. [7] Thostenson, E. T., Ren, Z., and Chou, T.-W., 2001, Advances in the Science and Technology of Carbon Nanotubes and Their Composites: A Review, Compos. Sci. Technol., 61(13), pp. 18991912. [8] Stankovich, S., Dikin, D. A., Dommett, G. H. B., Kohlhaas, K. M., Zimney, E. J., Stach, E. A., Piner, R. D., Nguyen, S. T., and Ruoff, R. S., 2006, Graphene-Based Composite Materials, Nature (London), 442(7100), pp. 282286. [9] McAllister, M. J., Li, J.-l., Adamson, D. H., Schniepp, H. C., Abdala, A. A., Liu, J., Herrera-Alonso, M., Milius, D. L., Car, R., Prudhomme, R. K., and Aksay, I. A., 2007, Single Sheet Functionalized Graphene by Oxidation and Thermal Expansion of Graphite, Chem. Mater., 19(18), pp. 43964404. [10] Raee, M. A., Raee, J., Wang, Z., Song, H., Yu, Z.-Z., and Koratkar, N., 2009, Enhanced Mechanical Properties of Nanocomposites at Low Graphene Content, ACS Nano, 3(12), pp. 38843890. [11] Ramanathan, T., Abdala, A. A., Stankovich, S., Dikin, D. A., Herrera-Alonso, M., Piner, R. D., Adamson, D. H., Schniepp, H. C., Chen, X., Ruoff, R. S., Nguyen, S. T., Aksay, I. A., Prudhomme, R. K., and Brinson, L. C., 2008, Functionalized Graphene Sheets for Polymer Nanocomposites, Nat. Nanotechnol., 3(6), pp. 327331. [12] Raee, M. A., Raee, J., Srivastava, I., Wang, Z., Song, H., Yu, Z.-Z., and Koratkar, N., 2010, Fracture and Fatigue in Graphene Nanocomposites, Small, 6(2), pp. 179183. [13] Schniepp, H. C., Li, J.-L., McAllister, M. J., Sai, H., Herrera-Alonso, M., Adamson, D. H., Prudhomme, R. K., Car, R., Saville, D. A., and Aksay, I. A., 2006, Functionalized Single Graphene Sheets Derived From Splitting Graphite Oxide, J. Phys. Chem. B, 110(17), pp. 85358539. [14] Zandiatashbar, A., Picu, C. R., and Koratkar, N., 2012, Control of Epoxy Creep Using Graphene, Small, in press. DOI: 10.1002/smll.201102686. [15] Shames, I. H., and Cozzarelli, F. A., 1997, Elastic and Inelastic Stress Analysis: Revised Printing, Taylor & Francis, Washington, D.C. [16] Zhang, W., Picu, R. C., and Koratkar, N., 2007, Suppression of Fatigue Crack Growth in Carbon Nanotube Composites, Appl. Phys. Lett., 91(19), p. 193109. [17] Erdogan, F., and Joseph, P. F., 1989, Toughening of Ceramics Through Crack Bridging by Ductile Particles, J. Am. Ceram. Soc., 72(2), pp. 262270. [18] Zhang, W., Picu, R. C., and Koratkar, N., 2008, The Effect of Carbon Nanotube Dimensions and Dispersion on the Fatigue Behavior of Epoxy Nanocomposites, Nanotechnology, 19(28), p. 285709. [19] Putz, K. W., Palmeri, M. J., Cohn, R. B., Andrews, R., and Brinson, L. C., 2008, Effect of Cross-Link Density on Interphase Creation in Polymer Nanocomposites, Macromolecules, 41(18), pp. 67526756. [20] Zandiatashbar, A., Picu, C. R., and Koratkar, N., 2011, Depth Sensing Indentation of Nanoscale Graphene Platelets in Nanocomposite Thin Films, Mater. Res. Soc. Symp. Proc., 1312, pp. 16.

Conclusions

A comprehensive analysis of the mechanical behavior of epoxyGPL nanocomposites is presented. The addition of nanosized GPL increases the creep resistance and slows down fatigue crack growth. The ow stress of the composite subjected to uniaxial monotonic loading is larger than that of neat epoxy, and the difference increases as the temperature increases. The strain hardening rate of both neat and lled epoxy increases with increasing temperature. The local mechanical behavior is probed, and it is concluded that GPL do not produce an interphase with modied mechanical properties, as usually seen in thermoplastics with nanoller additions. This suggests that the improved properties observed especially in the fracture behavior are primarily due to the interaction of the crack front with individual llers or to modied plasticity in the vicinity of the crack tip and not to the percolation of interphase zones.

031011-6 / Vol. 134, JULY 2012

Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 04/25/2013 Terms of Use: http://asme.org/terms

Vous aimerez peut-être aussi