Sie sind auf Seite 1von 7

Materials Science and Engineering A 430 (2006) 2733

Microstructures and tensile behavior of carbon nanotube reinforced Cu matrix nanocomposites


Kyung Tae Kim a , Seung Il Cha a , Seong Hyeon Hong b , Soon Hyung Hong a,
Department of Materials Science and Engineering, Korea Advanced Institute of Science and Technology, 373-1 Kusung-dong Yusung-gu, Daejeon 305-701, Republic of Korea b Nano P/M Group, Korea Institute of Machinery and Materials, 66 Sangnam-dong, Changwon, Kyungnam 641-010, Republic of Korea Received 25 January 2006; received in revised form 6 April 2006; accepted 27 April 2006
a

Abstract Carbon nanotubes (CNTs) have been considered as an ideal reinforcement to improve the mechanical performance of monolithic materials. However, the CNT/metal nanocomposites have shown lower strength than expected. In this study, the CNT reinforced Cu matrix nanocomposites were fabricated by spark plasma sintering (SPS) of high energy ball-milled nano-sized Cu powders with multi-wall CNTs, and followed by cold rolling process. The microstructure of CNT/Cu nanocomposites consists of two regions including CNT/Cu composite region, where most CNTs are distributed, and CNT free Cu matrix region. The stressstrain curves of CNT/Cu nanocomposites show a two-step yielding behavior, which is caused from the microstructural characteristics consisting of two regions and the load transfer between these regions. The CNT/Cu nanocomposites show a tensile strength of 281 MPa, which is approximately 1.6 times higher than that of monolithic Cu. It is conrmed that the key issue to enhance the strength of CNT/metal nanocomposite is homogeneous distribution of CNTs. 2006 Elsevier B.V. All rights reserved.
Keywords: Carbon nanotube; Cu; Nanocomposites; Microstructures; Two-step yielding

1. Introduction Carbon nanotubes have been expected as an ideal reinforcement to improve the mechanical performance of monolithic materials due to their high elastic modulus, strength and aspect ratio [1]. Recently, several researches have shown that the addition of carbon nanotube can considerably enhance the strength, toughness and conductivities of polymers and ceramics [29]. In CNT/polymer nanocomposite, the addition of carbon nanotube as reinforcement improves the tensile strength of the polymer matrix by several times [2]. In case of CNT/silica nanocomposites fabricated by solgel process, the bending strength and fracture toughness increases by about two and three times, respectively, compared to monolithic silica [9]. Also, CNT/alumina nanocomposite processed by molecular level mixing shows strengthening and toughening by addition of CNTs [10]. However, CNT/metal nanocomposites have shown inferior mechanical properties than expected compared to CNT/polymer

Corresponding author. Tel.: +82 42 869 3327; fax: +82 42 869 3310. E-mail address: shhong@kaist.ac.kr (S.H. Hong).

or CNT/ceramic nanocomposites [1115]. The inferior mechanical properties of CNT/metal nanocomposites are mainly due to severe agglomeration of CNTs and low relative densities, ranged 8595%, which is caused by conventional powder metallurgy process, consisting of mixing of CNTs with metal powders and followed by sintering or hot pressing process [1115]. Especially Zhan et al. [4] demonstrated that the dispersion and homogenous mixing between CNT and ceramic matrix could be obtained by mixing nano-sized matrix powders with CNTs. They showed that the SPS process is very promising technique for full densication of CNT/ceramic nanocomposites, separately from no reliable mechanical data refuted by Wang et al. [7]. Also, we developed the CNT/Cu nanocomposites, which show remarkable enhancement of yield strength compared to that of unreinforced Cu [16], there have been little attempts to improve CNT/metal nanocomposites by modication of traditional powder metallurgy process and to analyze the strengthening mechanism of CNT/metal nanocomposites. In this study, the microstructure and tensile behavior of CNT/Cu nanocomposites, fabricated with nano-sized Cu powders by spark plasma sintering process, were investigated. The tensile deformation behavior of CNT/Cu nanocomposites was analyzed based on the

0921-5093/$ see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2006.04.085

28

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733

load transfer theory, which is main strengthening mechanism of metal matrix composites. 2. Experimental procedures Nano-sized Cu powders were fabricated by spray drying process of [Cu(NO3 )2 ]3H2 O followed by burn-up and reduction processes. The water solution of [Cu(NO3 )2 ]3H2 O was spray dried on hot wall with 15,000 rpm. The dried powder was burnup at 300 C to fabricate oxide powder. The oxide powder was reduced into Cu powder at 200 C in H2 atmosphere. The powder size of Cu powders were ranged from 200 to 300 nm [17]. Multi-walled carbon nanotubes produced by a chemical vapor deposition (CVD) process, with an average diameter of about 40 nm and average length of a few m, were supplied from Iljin Nanotech Co., Ltd. The density of CNT used in this study is calculated by 1.8 g/cm3 from the TEM observation and the density of Cu is 8.9 g/cm3 . In this case, the weight percent of CNT is 1.0 for 5 vol.% of CNT/Cu nanocomposite and 2.2 for 10 vol.% of CNT/Cu nanocomposites. The nano-sized Cu powder and CNTs were mixed into composite powder through high energy ball milling process using planetary miller (Fritsch GmbH) for 24 h with 150 rpm. The volume fraction of CNT varies from 0 to 10%. The CNT/Cu composite powders were pre-compacted in a graphite mold under a pressure of 10 MPa. The pre-compacted powders were sintered by spark plasma sintering system at temperature of 700 C for 1 min in vacuum of 103 torr under a pressure of 50 MPa. The heating rate was maintained at 100 C/min. The spark plasma sintered CNT/Cu nanocomposites were cold rolled up to 50% reduction and followed by full annealing at 650 C for 3 h. The microstructures of CNT/Cu composite powders and sintered nanocomposites were analyzed by optical microscopy, high-resolution scanning electron microscopy (HRSEM). The volume fractions of CNTs in CNT/Cu nanocomposites were determined by analyzing the carbon contents by Elemental Analyzer (EA1110-FISONS) and by C/S analyzer (ELTRA CS800). Tensile tests were performed using INSTRON 5583 under a crosshead speed of 0.2 mm/min. Dog-bone shaped sub-size specimens with gage length of 9 mm and width of 2.5 mm based on the ASTM E8M were used for tensile test. The nano-scale hardness in CNT/Cu nanocomposite was measured by nanoindentation test (Nanoindentation XP). 3. Results and discussions 3.1. Microstructure of CNT/Cu nanocomposites The CNT/Cu composite powders fabricated by high energy ball milling process show homogeneously implanted CNTs on the surface of Cu powder, which has about 10 m, as shown in Fig. 1. Therefore, the agglomeration of CNTs can be avoided by high energy ball-milling process of nano-sized Cu and CNTs. The CNT/Cu composite powders, consisting of carbon nanotubes implanted on the surface of Cu powders, were consolidated by spark plasma sintering process to fabricate CNT/Cu nanocomposites. The relative densities of CNT/Cu nanocomposites after spark plasma sintering of composite powders were

Fig. 1. SEM micrographs of CNT/Cu composite powders. (a) SEM images showing carbon nanotubes are distributed in Cu powders and (b) SEM images showing the carbon nanotubes are embedded and implanted homogeneously in a Cu powder.

measured as 9798.5% regardless of CNT volume fraction as shown in Table 1. The relative density of CNT/Cu nanocomposites can be enhanced by cold rolling and annealing process above 99%. The microstructure of CNT/Cu nanocomposites, fabricated by high energy ball milling and spark plasma sintering process, consists of two regions as shown in Fig. 2(a and b). One is CNT free matrix region where the CNTs are not distributed. The other is CNT/Cu composite region around the matrix region where the most CNTs are mixed with Cu matrix as shown in Fig. 3. The two-region microstructure of spark plasma sintered CNT/Cu nanocomposites is caused by the distribution of CNTs in CNT/Cu composite powders formed during high energy ballmilling process. In CNT/Cu composite powders, the CNTs are implanted on the surface of Cu powder and form CNT/Cu composite at the surface of Cu powders as shown in Fig. 1. Therefore, after spark plasma sintering process, the surfaces of Cu powders, where the CNTs are implanted, induce CNT/Cu composite region and the inside of Cu powder, where the CNTs are not distributed, becomes CNT free matrix region.

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733 Table 1 Theoretical and measured densities of Cu and CNT/Cu nanocomposites after spark plasma sintering and cold rolling process Specimen As-sintered Measured density Cu 5 vol.% CNT/Cu 10 vol.% CNT/Cu 8.74 8.42 8.13 (g/cm3 ) Theoretical density 8.96 8.59 8.22 (g/cm3 ) Relative density (%) 97.5 98.5 98.1 As-rolled

29

Relative density (%) 99.0 99.3 99.1

The shape of CNT/Cu composite region is elongated and aligned so that the longitudinal direction becomes parallel to the rolling direction by cold rolling and followed by annealing process as shown in Fig. 2(c and d). Although the shape of CNT/Cu composite region looks similar to elongated cubic ber rather than cylindrical short ber, the critical aspect ratio between the former and the latter shows negligible differences. Thus, in this study to easily treat the mechanical properties of CNT/Cu nanocomposites, the rolled CNT/Cu composite region is considered as cylindrical short ber. 3.2. Tensile behavior of CNT/Cu nanocomposites The stressstrain curves are obtained from the tensile test of CNT/Cu nanocomposites as shown in Fig. 4(a), when the

loading direction is parallel to the rolling direction of specimen. In addition, the yield strength, tensile strength and Youngs modulus were measured from stressstrain curves as shown in Fig. 4(b). The tensile strength and Youngs modulus of CNT/Cu nanocmposites are increased with increasing the CNT volume fraction. The tensile strength of CNT/Cu nanocomposite with 10 vol.% CNT is measured as 281 MPa, which is approximately 1.6 times higher than that of unreinforced Cu fabricated by spark plasma sintering, cold rolling and annealing process. Youngs modulus of CNT/Cu nanocomposite with 10 vol.% CNT is measured as 137 GPa, which is almost two times higher than that of monolithic Cu, which is measured as 70 GPa. The yield strength increases from 135 to 197 MPa by addition of 10 vol.% CNT. The increased Youngs modulus and tensile strength indicates that the CNT/Cu nanocomposites are successfully fabricated from CNT and nano-sized Cu powder by high energy ball milling, spark

Fig. 2. Optical micrographs of CNT/Cu nanocomposites. (a) Spark plasma sintered 5 vol.% CNT/Cu nanocomposite, (b) spark plasma sintered 10 vol.% CNT/Cu nanocomposite, (c) cold rolled and annealed 5 vol.% CNT/Cu nanocomposite, (d) cold rolled and annealed 10 vol.% CNT/Cu nanocomposite.

30

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733

Fig. 3. SEM micrographs showing CNT/Cu composite region in 5 vol.% CNT/Cu nanocomposites where the CNTs are homogeneously mixed with Cu matrix.

plasma sintering and followed by cold rolling and annealing process. The two-step yielding behavior was observed during tensile test of CNT/Cu nanocomposites as shown in Fig. 4(a). It is expected that the two-step yielding behavior is caused from two-region microstructure of CNT/Cu nanocomposite. The
Fig. 5. Microstructural modeling of CNT/Cu nanocomposites. (a) 3Dimensional micrographs of CNT/Cu nanocomposites observed by SEM, (b) schematic microstructural modeling of a brous CNT/Cu composite region consisting of CNTs and Cu matrix and (c) schematic microstructural modeling of CNT/Cu nanocomposite consisting of brous CNT/Cu composite region and matrix region.

Fig. 4. (a) Stressstrain curves of CNT/Cu nanocomposites obtained by tensile test and (b) the variation of elastic modulus, yield strength and tensile strength of CNT/Cu nanocomposites with varying the volume fraction of CNTs.

microstructure of CNT/Cu nanocomposites shows that the carbon nanotubes are not homogeneously distributed in Cu matrix, but the carbon nanotubes are densely distributed in localized region, which is aligned parallel to the longitudinal direction as shown in Fig. 5(a). As a simplied model for microstructure of CNT/Cu nanocomposite, it is assumed that the microstructure consists of two parts, i.e. brous CNT/Cu composite region and matrix region. A brous CNT/Cu composite region can be assumed as a short ber, consisting of carbon nanotubes and Cu matrix, which have averaged properties of carbon nanotubes and Cu matrix. Then, the microstructure of CNT/Cu nanocomposites can be simply modeled as a short ber reinforced metal matrix composite as shown in Fig. 5(c). In order to estimate the yield strength of brous CNT/Cu composite region and Cu matrix region, hardness of each region was measured by nanoindentation test. The hardness values obtained from nanoindentation test were converted into yield strengths based on Weertmans results, which reported that the yield strength of Cu is equal to one third of the hardness measured by nanoindentation test [18]. From the measurement of hardness, yield strengths of brous CNT/Cu composite regions in 5 and 10 vol.% CNT/Cu nanocomposites are estimated about 370 and 589 MPa, respectively, while the yield strength of matrix region is estimated as 190 MPa as shown in Table 2. These results conrm that the carbon nanotubes are effective strengthening

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733 Table 2 The hardness and calculated yield strength of brous CNT/Cu region and Cu matrix region in CNT/Cu nanocomposites from nanoindentation test Region Measured hardness at 400 nm of indenting depth (GPa) Yield strength calculated from hardness (MPa) Cu matrix region 0.57 190 CNT/Cu composite region of 5 vol.% CNT/Cu nanocomposite 1.11 370

31

CNT/Cu composite region of 10 vol.% CNT/Cu nanocomposite 1.75 589

reinforcements when distributed homogeneously in metallic matrix. The stressstrain curve for tensile test can be divided into three different stages as shown in Fig. 6. It resembles the tensile behavior of ductile ber reinforced composite, where the bers undergo some plastic deformation before fracture [19]. Then the tensile behavior can be analyzed based on the shear-lag model, which is derived from the load transfer theory between matrix and bers in composite material [20]. The two-step yielding behavior of CNT/Cu nanocomposite is related to a difference in yield strengths between brous CNT/Cu composite region and matrix region. In the stage I of stressstrain curve in Fig. 6, both matrix region and composite region deform elastically up to primary yield strength. When the stress reaches the primary yield strength, the matrix region starts to deform plastically, while the composite region is still under elastic deformation because the yield strength of brous CNT/Cu composite region is higher than that of matrix region. Then, the primary yield strength of CNT/Cu nanocomposite can be analyzed by assuming that the elastic bers of composite regions are distributed within plastic Cu matrix. The primary yield strength, y,1 , can be estimated by equation (1) based on the generalized shear-lag model [20] suggested for misaligned ber reinforced composite, y,1 = Vf m Seff + m 2 3 4 3 1+ 1 S sin2 (1)

where Seff is effective aspect ratio of a misaligned ber and is misorientation angle between longitudinal axis of brous CNT/Cu composite region and loading axis of tensile test. The effective aspect ratio in Eq. (2) is a function of misorientation angle of brous CNT/Cu composite region. Therefore, in order to obtain average effective aspect ratio of brous CNT/Cu composite region, the probability density function of misorientation angle of brous CNT/Cu composite region should be introduced as followed [20], Seff,a =
/2 0

Seff ( )F ( )(2 sin )d

(3)

where Seff,a is average effective aspect ratio and F is probability density function of misorientation angle of brous CNT/Cu

Seff = S cos2 +

(2)

Fig. 6. Stressstrain curve showing two-step yielding phenomena, which is dependent on the three different stages (IIII) of deformation, during tensile test of 10 vol.% CNT/Cu nanocomposite.

Fig. 7. The probability density function of (a) 5 vol.% and (b) 10 vol.% CNT reinforced Cu matrix nanocomposites from the image analysis of microstructure.

32

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733

Table 3 Comparison of primary and secondary yield strengths of CNT/Cu nanocomposites measured by tensile test and calculated by generalized shear-lag model Volume fraction of CNTs (%) 5 10 Volume fraction of CNT/Cu composite region (%) 6.3 12.7 Effective aspect ratio of CNT/Cu composite region (Seff ) 3.2 4.8 Measured primary yield strength (MPa) 149 197 Calculated primary yield strength (MPa) 151 180 Measured secondary yield strength (MPa) 180 240 Calculated secondary yield strength (MPa) 167 256

composite region in Eq. (3). The probability density function of misorientation angle of brous CNT/Cu composite region is measured by image analysis of microstructure as shown in Fig. 7. Then the average effective aspect ratio of brous CNT/Cu composite region and primary yield strength of CNT/Cu nanocomposite is calculated as shown in Table 3. In Table 3, the calculated primary yield strengths of CNT/Cu nanocomposite based on the generalized shear-lag model show good agreement with those measured from the tensile test, when 5 and 10 vol.% of CNTs are reinforced. In the stage II of stressstrain curve in Fig. 6, the matrix region deforms plastically but the brous CNT/Cu composite region is still under elastic deformation. However, if the stress is increased further, before the brous CNT/Cu composite region start to deform plastically, work hardening of unreinforced Cu matrix slightly in advance is generated in low strain region around 1%. Then when the load transferred from matrix to composite region increases above the yield strength of composite region above 1% strain. Then the secondary yielding is observed when the stress transferred from matrix region to composite region exceeds the yield strength of composite region. It is expected that the CNTs in CNT/Cu composite region are randomly layered into the slab shapes on the Cu particles during ball milling process. Therefore, rule-of-mixture model rather than shear-lag model is much more appropriate to analyze the yielding behavior of CNT/Cu composite region itself. Therefore, the secondary yield strength of CNT/Cu nanocomposites, y,2 , related to the yielding of CNT/Cu composite region, can be estimated based on the rule-of-mixture expressed as Eq. (4), assuming that work hardening of Cu matrix is negligible between the primary yield strength and the secondary yield strength, y,2 = y,1 (1 Vf ) + f Vf (4)

4. Conclusion CNT/Cu nanocomposites have been fabricated by spark plasma sintering of high energy ball-milled nano-sized Cu powders with multi-wall CNTs, and followed by cold rolling process. The microstructure of CNT/Cu nanocomposite can be modeled as two-region structure, consisting of brous CNT/Cu composite region and CNT free matrix region. Due to this characteristic inhomogeneous distribution of CNTs in Cu matrix, the mechanical properties of CNT/Cu nanocomposites show a two-step yielding behavior in stressstrain curve. The measured primary and secondary yield strength of CNT/Cu nanocomposites are reasonably agreed with estimated values, based on the generalized shear-lag model. It shows that the strengthening mechanism is mainly caused by a considerable load transfer of elastically stiff CNTs in plastic matrix. The homogeneous distribution of CNTs in brous CNT/Cu composite region shows considerably improved yield strength. These results indicate that the homogeneous distribution of CNTs in metal matrix is the most critical issues to enhance the mechanical properties of CNT/metal nanocomposites. Acknowledgements This research was supported by a grant (code #06K150100510) from Center for Nanostructured Materials Technology under 21st Century Frontier R&D Programs of the Ministry of Science and Technology, Korea. References
[1] P.J.F. Harris, Int. Mater. Rev. 49 (1) (2004) 3143. [2] A.A. Mamedov, N.A. Kotov, M. Prato, D.M. Guldi, J.P. Wicksted, A. Hirsch, Nat. Mater. 1 (2002) 15. [3] H. Koerner, G. Price, N.A. Pearce, M. Alexander, R.A. Vaia, Nat. Mater. 3 (2004) 115120. [4] G.-D. Zhan, J.D. Kuntz, J. Wan, A.K. Murkherjee, Nat. Mater. 2 (2003) 3842. [5] A. Peigney, E. Flahaut, Ch. Laurent, F. Chastel, A. Rosset, Chem. Phys. Lett. 352 (2002) 2025. [6] E. Flahaut, A. Peigney, Ch. Laurent, Ch. Marliere, F. Chastel, A. Rosset, Acta Mater. 48 (2000) 38033812. [7] X. Wang, N.P. Padture, H. Tanaka, Nat. Mater. 3 (2004) 539544. [8] K.T. Lau, S.-Q. Shi, H.-M. Cheng, Comp. Sci. Tech. 63 (2003) 11611164. [9] J. Ning, J. Zhang, Y. Pan, J. Guo, Mater. Sci. Eng. A357 (2003) 392 396. [10] S.I. Cha, K.T. Kim, K.H. Lee, C.B. Mo, S.H. Hong, Scr. Mater. 53 (2005) 793797.

where Vf is volume fraction and f is yield strengths of brous CNT/Cu composite region. Both the measured and calculated secondary yield strength show good agreement each other as compared in Table 3. In stage III of stressstrain curve in Fig. 6, both matrix region and composite region deform plastically as the stress exceeds the secondary yield strength of CNT/Cu nanocomposite. From these analyses, it is shown that the twostep yielding behavior of CNT/Cu nanocomposite is caused from two-region microstructure and from the load transfer between two regions, which can be reasonably predicted by the generalized shear-lag model.

K.T. Kim et al. / Materials Science and Engineering A 430 (2006) 2733 [11] S.R. Dong, J.P. Tu, X.B. Zhang, Mater. Sci. Eng. A313 (2001) 8387. [12] T. Kuzumaki, K. Miyazawa, H. Ichinose, K. Ito, J. Mater. Res. 13 (1998) 24452449. [13] C.L. Xu, B.Q. Wei, R.Z. Ma, J. Liang, X.K. Ma, D.H. Wu, Carbon 37 (1999) 855858. [14] R. Zhong, H. Cong, P. Hou, Carbon 41 (2003) CO1CO851. [15] L.Y. Wang, J.P. Tu, W.X. Chen, Y.C. Wang, X.K. Liu, C. Olk, D.H. Cheng, X.B. Zhang, Wear 254 (2003) 12891293.

33

[16] S.I. Cha, K.T. Kim, S.N. Arshad, C.B. Mo, S.H. Hong, Adv. Mater. 17 (2005) 13771381. [17] S.-H. Hong, B.-K. Kim, Mater. Lett. 4377 (2003) 17. [18] J. Weertman, D. Farkas, K. Hemker, H. Kung, M. Mayo, R. Mitra, H.V. Swygenhoven, MRS Bull. 24 (2) (1999) 4450. [19] T.H. Courtney, Mechanical Behavior of Materials, McGraw-Hill Book Co., 1990, pp. 224231. [20] H.J. Ryu, S.I. Cha, S.H. Hong, J. Mater. Res. 18 (2003) 28512853.

Das könnte Ihnen auch gefallen