Electrical conductivity phenomena in an epoxy resin–carbon-based materials composite

August 20, 2017 | Autor: Pravin Jagdale | Categoría: Materials Engineering, Mechanical Engineering, Aerospace Engineering
Share Embed


Descripción

Composites: Part A 61 (2014) 108–114

Contents lists available at ScienceDirect

Composites: Part A journal homepage: www.elsevier.com/locate/compositesa

Electrical conductivity phenomena in an epoxy resin–carbon-based materials composite Castellino Micaela a,⇑, Chiolerio Alessandro a, Shahzad Muhammad Imran b, Jagdale Pravin Vitthal b, Tagliaferro Alberto b a b

Center for Space Human Robotics, Istituto Italiano di Tecnologia, Corso Trento 21, Torino 10129, Italy Applied Science and Technology Department, Politecnico di Torino, Corso Duca degli Abruzzi 24, Torino 10129, Italy

a r t i c l e

i n f o

Article history: Received 16 January 2013 Received in revised form 7 February 2014 Accepted 10 February 2014 Available online 20 February 2014 Keywords: A. Polymer–matrix composites (PMCs) B. Electrical properties C. Finite element analysis (FEA) D. Physical methods of analysis

a b s t r a c t Nanocomposites (NCs) were prepared using a thermoset commercial epoxy resin and a variety of carbonbased materials (CBMs): carbon beads and powders and 13 different commercial Carbon NanoTubes (CNTs), with the aim of comparing their performance. Electrical behaviour of NCs prepared with two CBMs weight concentrations (1 and 3 wt.%) was investigated. The more promising were selected to study their conductivity behaviour more in detail (5 different concentrations in the range 0–5 wt.%). A thorough electrical characterization, performed with the support of a Finite Element Method (FEM) simulation, allowed us to estimate the resistivities and to apply physical models such as percolation and fluctuation-mediated tunnelling theories. Several conduction behaviours have been found following CBMs type used and its concentration: highly conductive NCs, non-linear diode-like trend and highly resistive. Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction The development of novel functional materials and of sustainable processes, that lead to products with improved performance, is an open challenge. In fact, the design of materials with new properties for specific applications is the key element in the creation of high value-added products in multidisciplinary technological areas, such as bioscience, optoelectronics, nano-electronics and nano-photonics. In this context, polymer-based composites cover applications ranging from biodegradable packaging, with reduced environmental impact, to biomedical implants [1,2]. The exploration of the nano-world has allowed the integration of nanostructures into polymer matrices with hierarchical architectures [3]. In this way either by compatibilization of nanofillers with different properties or by controlling the collective communication between particles throughout an organized network, it is possible to tailor dielectric, electrical, mechanical, thermal and magnetic properties of nanocomposites [4–9]. In fact, the nanofillers affinity with the polymeric matrices leads to the appearance of synergic structural and ⇑ Corresponding author. Tel.: +39 011 564 7344; fax: +39 011 564 7399. E-mail addresses: [email protected] (M. Castellino), alessandro.chiolerio@ iit.it (A. Chiolerio), [email protected] (M.I. Shahzad), [email protected] (P.V. Jagdale), [email protected] (A. Tagliaferro). http://dx.doi.org/10.1016/j.compositesa.2014.02.012 1359-835X/Ó 2014 Elsevier Ltd. All rights reserved.

chemical–physical phenomenon not directly attributable to the individual non-interacting components [10]. It has to be pointed out that nanofillers, which provide large interfacial areas, due to their intrinsic high surface-volume ratio, are suitable to create high density composites with a reduced percolation threshold [11]. As a result, the transfer of the properties from the nanoscale to macroscopic materials is very efficient also at low loads (10 >10

>95.0 >95.0 >95.0 >90.0 >90.0 >70.0 >70.0 >95.0 >70.0 >95.0 >60.0 >95.0 >99.0 97.0 98.5 >90.0

110

M. Castellino et al. / Composites: Part A 61 (2014) 108–114

Fig. 1. FESEM images of CBM #9, #10 and #11, as prepared.

Fig. 2. Composites made with epoxy resin with CBM #8 at different wt.% (from 0 to 5 wt.%). The round spot on each side is due to the silver paste, used as electrodes for the IV measurements.

by the commercial code, setting a fine quality of the elements and a progressive element growth, from a minimum element size corresponding to one hundredth of the smallest dimension (3 mm). Boundary conditions are such that a finite potential is kept on the electrodes (either zero or the positive maximum voltage provided by the DC source) and all other surfaces are kept in an ‘‘insulation’’ condition with respect to external volume, hence having no current density component parallel to the surface normal vector. Solution time for each resistivity entry is over 300 s (Intel™ CoreÒ 2 Quad Q9550 2.83 GHz 4 GB DDR3). Convergence is reached when the error (relative difference in the current values between two subsequent minimization steps of FEM code) falls below 106. 2.3. Electrical characterization Electrical measurements were performed using the so called ‘‘Two Point Probe (TPP) method’’ [41] with a Keithley-238 high current source measure unit, used as high voltage source and nanoamperometer. 3. Results and discussion FEM simulations allowed us to estimate the correction factor to be applied to the geometry in order to properly take into account

the real flow of current in the calculation of the resistivity values. An example of the simulation control volume is given in Fig. 4, where two electrodes, having a diameter of 5 mm each, have been placed at the edges of the sample, on front and on rear surface respectively. The current density is distributed almost in the whole sample, with the exception of the portions close to the electrodes, where the effective paths avoid the sample bottom and edges. Based on these simulations, the effective electrical path was estimated to be of thickness 3 mm (same as the sample), width 3 cm (same as the sample) and length 1 cm (sample length reduced due to electrodes size and dead ends). Another electrode arrangement was tested in order to verify the solidity of our approach. When the electrodes are facing one each other on top and bottom sides (see Fig. 5), the volume, in which a non-marginal current density flows, is approximately corresponding to a cylinder having a diameter twice that of the real electrode. In order to find out if the electrodes displacement could affect our results, before testing all the composites, we performed current–voltage (IV) measurements on one selected composite (epoxy + CBM #8) at 3 wt.% with both geometries, finding out that the electrodes configurations do not lead to substantial differences in the conduction behaviour. In fact we obtained similar values for resistivities (39.4 X cm and 36.1 X cm respectively, deviation within 8%). We eventually decided to use the front–back diagonal layout (Fig. 4) for the

M. Castellino et al. / Composites: Part A 61 (2014) 108–114

111

Fig. 3. FESEM images of epoxy composites made with (a) CBM #5 at 3 wt.%; (b) CBM #10 at 3 wt.%; (c) CBM #13 at 1 wt.% and (d) CBM #16 at 3 wt.%.

Fig. 4. Current density simulation with front–back diagonal electrodes geometry.

comparison of NCs, since (i) it involves almost the entire sample volume in the current flow and (ii) it is more easily handled during IV measurements. In Fig. 6 we have reported all the IV curves measured for each sample at 1 and 3 wt.% for all the 16 composites. Most of the samples with 1 wt.% of CBMs show a resistivity (q) value >109 X cm with very noisy signals (see black curves in Fig. 6) that can be considered as typical of insulator materials. However samples #3 and #8, with the same amount of fillers, have shown already remarkable q values in the range 102–103 X cm, with an almost but not fully linear response in the voltage range ±10 V. Increasing the amount of #3 and #8 CBMs to 3 wt.%, resulted in a perfect linear response with q values in the range 100–101 X cm (red curves), while the other NCs, with the same amount of fillers, decreased their resistivity values only down to 105–106 X cm with almost linear curves. We want to stress that #3 and #8 CBMs are the same

kind of MWCNTs (produced by the same company), the first as grown and the latter lightly functionalized with COOH groups. They differ from the others CBMs mainly because of their average length (1.5 lm), which are the shortest among all the MWCNTs (see Table 1). So it seems that the functionalization does not substantially increase or reduce the electrical conductivity of the filler, while it has been proved that the thermal properties can be affected by a functionalization process [42]. SWCNTs, independently from their lengths and functionalization, lead to worse results probably due to the presence of clustering during the dispersion procedure, (see Fig. 3a), which lowers the efficiency of the filler. CBMs #9, #10 and #11, which have been produced in our lab, show the worst behaviour among all, since their IV curves point out that their resistivity values remain almost constant and indeed high (>1010 X cm), even at 3 wt.%. This can be justified by the fact that spherical and oblate fillers (CBM #10), due to their geometry,

112

M. Castellino et al. / Composites: Part A 61 (2014) 108–114

Fig. 5. Current density simulation with front–back vertical electrodes geometry.

Fig. 6. IV curves for all the samples prepared with 1 (black curve) and 3 (red curve) wt.% of CBMs. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

M. Castellino et al. / Composites: Part A 61 (2014) 108–114

113

according to this model is more accurate than the previous one, although still not fully compliant with data. So, after an exhaustive search in literature, we have found a revised tunnelling–percolation theory by Ambrosetti et al. [43,50], where they have extended the study of the conductivity dependencies by applying the low-density hopping like approach to higher densities. As it can be seen in Fig. 7 (red dots line) this time the curve fitting is approaching the experimental data in a better way, apart from the lowest measurable conductivity value (0 wt.%) which is limited by the experimental setup or by the polymer matrix intrinsic conductivity. 4. Conclusions

Fig. 7. Epoxy + CBM #8 conductivity as a function of dispersoid volume fraction fitted by percolation [46], tunnelling [49] and a revised tunnelling model [43].

are able to create percolating network if their amount (wt.%) is higher compared with cylindrical and prolate one [43]. On the other hand, CBM #9 and #11, which are made up by a mixture of highly entangled carbon structures (MWCNTs) and amorphous structures, are not suitable to be chosen as good fillers. Also CBM #2 and #12, which are the same kind of MWCNTs produced by the same company, the first one as grown and the latter graphitized, show the same behaviour, highly insulating, both at 1 and 3 wt.%. In this case these MWCNTs possess small diameters ( 109 X). The best performances have been reached by the shortest and thinner MWCNTs (both as grown and slightly functionalized with COOH groups), which can underline that small fillers with a high aspect-ratio (L/D = 158) can be better dispersed inside the composite and create a better percolating network within the matrix. Moreover, since all the CBMs have been dispersed in the same weight percentage, the smallest the dimensions the higher the concentration of fillers nanoparticles inside the composites. In fact, the relationship between the wt.% and the numbers of CNTs per unit volume is given by:

wt:% ¼

mCNT  NCNTs k

ð1Þ

where k is a constant depending on the total volume of the composite and the polymer matrix density, while the mass of a single CNT (mCNT) has been calculated according to the formula:

mCNT ¼ rm  p  L  ðNw  ðDext þ dw Þ  dw  ðNw Þ2 Þ

ð2Þ

where L is the length of a CNT, Dext and Dint are respectively the external and internal diameters expressed in nm, Nw is the approximate number of walls for MWCNTs, considering walls spaced by a distance dw = 0.345 nm (lattice spacing of graphite [51]) and rm = 7.6  104 g/m2 is a graphene sheet mass density [52]. We have applied physical models such as the percolation theory and the fluctuation-mediated tunnelling theory to the most conductive NCs, with poor agreement between experimental data and theoretical prediction. Then we have applied a recent revised tunnelling model obtaining a good fit result. Nevertheless a new conductivity model is needed, which has to take into account for real composite characteristics like the presence of impurities, structure and spatial distribution inside the polymer matrix together with defects and agglomeration problems [53]. Acknowledgment Funding has been supported by Politecnico di Torino and Fondi Regione Piemonte (DR 360/2008) within the project ‘‘Charge transport measurements in advanced materials: novel superconductors and nanostructured semiconductors for light harvesting’’. The authors wish to thank Dr. S. Guastella and Dr. A. Virga (Politecnico di Torino) for their FESEM measurements. References [1] Koo JH. Polymer nanocomposites: processing, characterization and applications. New York: McGraw-Hill; 2006. [2] Li C, Thostenson ET, Chou TW. Sensors and actuators based on carbon nanotubes and their composites: a review. Compos Sci Tech 2008;68:1227–49.

114

M. Castellino et al. / Composites: Part A 61 (2014) 108–114

[3] Park C, Yoon J, Thomas EL. Enabling nanotechnology with self assembled block copolymer patterns. Polymer 2003;44:6725–60. [4] Caseri W. Nanocomposites of polymers and metals or semiconductors: historical background and optical properties. Macromol Rapid Commun 2000;21:705–22. [5] Chiolerio A, Musso S, Sangermano M, Giorcelli M, Bianco S, Coisson M, et al. Preparation of polymer-based composite with magnetic anisotropy by oriented carbon nanotube dispersion. Diamond Relat Mater 2008;17:1590–5. [6] Bortz DR, Merino C, Martin-Gullon I. Carbon nanofibers enhance the fracture toughness and fatigue performance of a structural epoxy system. Compos Sci Tech 2011;71:31–8. [7] Martone A, Faiella G, Antonucci V, Giordano M, Zarrelli M. The effect of the aspect ratio of carbon nanotubes on their effective reinforcement modulus in an epoxy matrix. Compos Sci Tech 2011;71:1117–23. [8] Mora RJ, Vilatela JJ, Windle AH. Properties of composites of carbon nanotube fibres. Compos Sci Tech 2009;69:1558–63. [9] Biercuk MJ, Llaguno MC, Radosavljevic M, Hyun JK, Johnson AT, Fischer JE. Carbon nanotube composites for thermal management. Appl Phys Lett 2002;80:2767–9. [10] Hansen N, Adams DO, Fullwood DT. Quantitative methods for correlating dispersion and electrical conductivity in conductor–polymer nanostrand composites. Composites: Part A 2012;43:1939–46. [11] Gojny FH, Wichmann MHG, Fiedler B, Kinloch IA, Bauhofer W, Windle AH, et al. Evaluation and identification of electrical and thermal conduction mechanisms in carbon nanotube/epoxy composites. Polymer 2006;47: 2036–45. [12] Maldovan M, Ullal CK, Carter WC, Thomas EL. Exploring for 3D photonic bandgap structures in the 11 f.c.c. space groups. Nat Mater 2003;2:664–7. [13] Balberg I. Tunneling and nonuniversal conductivity in composite materials. Phys Rev Lett 1987;59:1305–8. [14] Byrne MT, Gun’ko YK. Recent advances in research on carbon nanotube– polymer composites. Adv Mater 2010;22:1672–88. [15] Suhr J, Victor P, Ci L, Sreekala S, Zhang X, Nalamasu O, et al. Fatigue resistance of aligned carbon nanotube arrays under cyclic compression. Nat Nanotechnol 2007;2:417–21. [16] Hone J, Llaguno MC, Nems NM, Johnson AT, Fisher JE, Walters DA, et al. Electrical and thermal transport properties of magnetically aligned single wall carbon nanotube films. Appl Phys Lett 2000;77:666–8. [17] Yu A, Itkis ME, Bekyarova E, Haddon RC. Effect of single-walled carbon nanotube purity on the thermal conductivity of carbon nanotube-based composites. Appl Phys Lett 2006;89:133102–4. [18] Sekitani T, Noguchi Y, Hata K, Fukushima T, Aida T, Someya T. A rubberlike stretchable active matrix using elastic conductors. Science 2008;321:1468–72. [19] Fujii M, Zhang X, Xie H, Ago H, Takahashi K, Ikuta T. Measuring the thermal conductivity of a single carbon nanotube. Phys Rev Lett 2005;95:0655021–24. [20] Kim P, Shi L, Majumdar A, McEuen PL. Thermal transport measurements of individual multiwalled nanotubes. Phys Rev Lett 2001;87:2155021–24. [21] Tsai MY, Yu CY, Yang CH, Tai NH, Perng TY, Tu CM, et al. Electrical transport properties of individual disordered multiwalled carbon nanotubes. Appl Phys Lett 2006;89:1921151–53. [22] Breuer O, Sundararaj U. Big returns from small fibers: a review of polymer/ carbon nanotube composites. Polym Comp 2004;25:630–45. [23] Pan Y, Li L, Chan SH, Zhao J. Correlation between dispersion state and electrical conductivity of MWCNTs/PP composites prepared by melt blending. Composites: Part A 2010;41:419–26. [24] MacDiarmid AG. Synthetic metals: a novel role for organic polymers. Synth Met 2002;125:11–22. [25] Villmow T, Pegel S, Pötschke P, Heinrich G. Polymer/carbon nanotube composites for liquid sensing: model for electrical response characteristics. Polymer 2011;52:2276–85. [26] Dresselhaus MS, Dresselhaus G, Charlier JC, Hernández E. Electronic, thermal and mechanical properties of carbon nanotubes. Philos Trans R Soc London, A 2004;362:2065–98. [27] Ma PC, Siddiqui NA, Marom G, Kim JK. Dispersion and functionalization of carbon nanotubes for polymer-based nanocomposites: a review. Composites: Part A 2010;41:1345–67. [28] Tasis D, Tagmatarchis N, Bianco A, Prato M. Chemistry of carbon nanotubes. Chem Rev 2006;106:1105–36. [29] Dalmas F, Dendievel R, Chazeau L, Cavaille JY, Gauthier C. Carbon nanotube– filled polymer composites. Numerical simulation of electrical conductivity in three-dimensional entangled fibrous networks. Acta Mater 2006;54:2923–31.

[30] Feng C, Jiang L. Micromechanics modeling of the electrical conductivity of carbon nanotube (CNT)–polymer nanocomposites. . [31] , , , , , , . [32] Wescott JT. Conductivity of carbon nanotube polymer composites. Appl Phys Lett 2007;90:033116. [33] Scocchi G, Ortona A, Grossi L, Bianchi G, D’Angelo C, Leidi T, Gilardi R. Evaluation of a simple finite element method for the calculation of effective electrical conductivity of compression moulded polymer–graphite composites. . [34] Rana S, Alagirusamy R, Joshi M. Development of carbon nanofibre incorporated three phase carbon/epoxy composites with enhanced mechanical, electrical and thermal properties. Composites: Part A 2011;42:439–45. [35] Liu CH, Fan SS. Nonlinear electrical conducting behavior of carbon nanotube networks in silicone elastomer. Appl Phys Lett 2007;90:041905–7. [36] Huang YY, Terentjev EM. Tailoring the electrical properties of carbon nanotube–polymer composites. Adv Funct Mater 2010;20:4062–8. [37] Li C, Thostenson ET, Chou TW. Dominant role of tunneling resistance in the electrical conductivity of carbon nanotube-based composites. Appl Phys Lett 2007;91:223114. [38] Tishkova V, Raynal PI, Puech P, Lonjon A, Le Fournier M, Demont P, et al. Electrical conductivity and Raman imaging of double wall carbon nanotubes in a polymer matrix. Compos Sci Tech 2011;71:1326–30. [39] Chiolerio A, Castellino M, Jagdale P, Giorcelli M, Bianco S, Tagliaferro A. Electrical properties of CNT-based polymeric matrix nanocomposites. In: Yellampalli Siva, editor. Carbon nanotubes-polymer nanocomposites. Rijeka, Croatia: INTECH Open Access Publisher; 2011. p. 215–30. [40] http://www.leuna-harze.de/products/epiloxreg-epoxy-resins/?L=1. [41] Schroder DK. Semiconductor material and device characterization. Hoboken: Wiley Inter-Science; 1990. [42] Gulotty R, Castellino M, Jagdale P, Tagliaferro A, Balandin A. Effects of functionalization on thermal properties of single-wall and multi-wall carbon nanotube–polymer nanocomposites. ACS Nano 2013;7:5114–21. [43] Ambrosetti G, Grimaldi C, Balberg I, Maeder T, Danani A, Ryser P. Solution of the tunneling–percolation problem in the nanocomposite regime. Phys Rev B 2010;81:155434. [44] Balberg I. The importance of bendability in the percolation behavior of carbon nanotube and graphene–polymer composites. J Appl Phys 2012;112:066104. [45] Coleman JN, Curran S, Dalton AB, Davey AP, McCarthy B, Blau W, et al. Percolation-dominated conductivity in a conjugated-polymer–carbonnanotube composite. Phys Rev B 1998;58:R7492–5. [46] Stauffer D, Aharony A. Introduction to percolation theory. London: Taylor and Francis; 1994. [47] Chiolerio A, Sangermano M. In situ synthesis of Ag-acrylic nanocomposites: tomography-based percolation model, irreversible photoinduced electromigration and reversible electromigration. Mater Sci Eng B 2012;177: 373–80. [48] Kilbride BE, Coleman JN, Fraysse J, Fournet P, Cadek M, Drury A, et al. Experimental observation of scaling laws for alternating current and direct current conductivity in polymer–carbon nanotube composite thin films. J Appl Phys 2002;92:4024–30. [49] Allaoui A, Hoa SV, Pugh MD. The electronic transport properties and microstructure of carbon nanofiber/epoxy composites. Compos Sci Tech 2008;68:410–6. [50] Ambrosetti G, Balberg I, Grimaldi C. Percolation-to-hopping crossover in conductor–insulator composites. Phys Rev B 2010;82:134201–7. [51] Osipov VY, Enoki T, Takai K, Takahara K, Endo M, Hayashi T, et al. Magnetic and high resolution TEM studies of nanographite derived from nanodiamond. Carbon 2006;44:1225–34. [52] Peigney A, Laurent Ch, Flahaut E, Bacsa RR, Rousset A. Specific surface area of carbon nanotubes and bundles of carbon nanotubes. Carbon 2001;39:507–14. [53] Hu N, Karube Y, Yan C, Masuda Z, Fukunaga H. Tunneling effect in a polymer/ carbon nanotube nanocomposite strain sensor. Acta Mater 2008;56:2929–36.

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.