You are on page 1of 24

Accepted Manuscript

Enhanced mechanical and electrical properties of novel graphene reinforced copper


matrix composites

C. Salvo, R.V. Mangalaraja, R. Udayabashkar, M. Lopez, C. Aguilar

PII: S0925-8388(18)34056-8
DOI: https://doi.org/10.1016/j.jallcom.2018.10.357
Reference: JALCOM 48175

To appear in: Journal of Alloys and Compounds

Received Date: 24 July 2018


Revised Date: 22 October 2018
Accepted Date: 27 October 2018

Please cite this article as: C. Salvo, R.V. Mangalaraja, R. Udayabashkar, M. Lopez, C. Aguilar,
Enhanced mechanical and electrical properties of novel graphene reinforced copper matrix composites,
Journal of Alloys and Compounds (2018), doi: https://doi.org/10.1016/j.jallcom.2018.10.357.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Enhanced mechanical and electrical properties of novel graphene reinforced copper matrix
composites

C. Salvo1*, R.V. Mangalaraja1*, R. Udayabashkar1, M. Lopez1 and C. Aguilar2

PT
1
Advanced Ceramics and Nanotechnology Laboratory, Department of Materials Engineering,
Faculty of Engineering, University of Concepcion, Concepcion, Chile.

RI
2
Departament of Metallurgical and Materials Engineering, Federico Santa María Technical
University, Valparaiso, Chile.

SC
*Corresponding Author: mangal@udec.cl, csalvo@udec.cl

Abstract
U
AN
By this report, we described the synthesis of graphene (1 wt%) reinforced copper

matrix composites (Cu-GNS) by using the mechanical milling and the hot pressing. The structure
M

and morphology of the few layer graphene nanosheets (GNS) reinforced copper matrix
D

composites were discussed with the X-ray diffraction (XRD) and scanning electron microscope
TE

(SEM) techniques, respectively. The influence of the mechanical milling on the microstructure

was analyzed. The chemical compatibility and successful formation of the composites were
EP

confirmed by the X-ray diffraction studies. The consolidation of the powders was performed by

using vacuum hot-press sintering (at 600 and 700°C). The hardness and electrical conductivity
C

were discussed with respect to the sintering temperatures. An improvement in the electrical
AC

conductivity around 22% for 1 wt% graphene reinforced Cu when compared with the pure Cu

matrix consolidated at 600oC was observed. The importance of lower sintering temperature and

higher applied load in the consolidation (600°C, 30 MPa for 30 min) to obtain the Cu-GNS

composites with improved physical, mechanical and electrical properties was emphasized.

Keywords: Cu-based composites; graphene nanosheets, electrical conductivity.


ACCEPTED MANUSCRIPT

1. Introduction

Copper (Cu) is widely used in high-end application areas such as integrated circuits,

electric switches and electronic packages, by virtue of its distinguishable properties such as high

electrical conductivity, fatigue resistance, workability and corrosion resistance. However, it has

PT
limited mechanical properties. The copper-based composites have been under investigation by

RI
various researchers [1-3] to address the requirements for their utility in the functional

applications. The metal matrix composites (MMC) reinforced with different particles exhibits a

SC
combination of properties of their constituents, shown improved properties when compared with

pure (unreinforced) matrix. The strengthening mechanisms involved in MMC have been

U
discussed previously by various research groups [4-6]. The key to fabricate materials with high
AN
mechanical and electrical properties is to produce a microstructure in which the dislocation

motions are blocked and scattering of conducting electrons is minimal [7]. In general, ex-situ or
M

in-situ methods [8] are used as different strategies for dispersing reinforcement particles in Cu
D

matrix to improve the mechanical properties. The quantity of reinforcement added to the metal
TE

matrix will significantly influence the microstructure and functional properties of the composites.

The required properties can be achieved by varying the amount of weight fraction of the
EP

reinforcement particles [9]. The homogeneous distribution of the reinforcement particles is the

most important factor to obtain the composites with improved mechanical properties [10]. The
C

particle size of the reinforcements has a strong effect on the mechanical properties of composites,
AC

thus the preference has been given to the nanosized particulates as reinforcing materials [6, 9].

Graphene, the allotrope of carbon, is a single to few layers of covalently bonded sp2

carbon atoms, hexagonally packed in a honeycomb crystal structure [10, 11] and was synthesized

effectively the first time by Geim and Novoselov in the year 2004 [12]. The layer structured

graphene nanosheet (GNS) highly attracted the researchers with its large surface area and high
ACCEPTED MANUSCRIPT

conductivity [13]. The unique electrical, mechanical and thermal properties of graphene made

this material as a promising reinforcement for Cu matrix [9, 13-21]. The addition of graphene to

the copper matrix can enhances the electrical properties due to the high charge-carrier mobility.

The ability to reduce effectively the movement of dislocations [22], also the effect of interfaces

PT
between graphene and metal matrix for having good bonding [23] are influenced by the

RI
agglomeration tendency and poor wettability of Cu and graphene. To address this problem, the

mechanical milling (ex-situ method) is considered as a most viable method to produce an uniform

SC
dispersion of reinforcement particles in the MMC [24, 25].

In this work, we report the realization of graphene nanosheets (GNS) reinforced copper

U
matrix composites by using the mechanical milling and subsequent hot pressed. After a detailed
AN
literature review, 1 wt. % of GNS as reinforcement is considered as the optimal parameter for the

current work. The consolidation of the powders was performed by using hot-press sintering at
M

600 and 700°C, respectively. The influence of milling time on microstructural morphology and
D

the effect of consolidating temperature on the mechanical and electrical properties of the 1 wt. %
TE

GNS reinforced Cu matrix were discussed. The current study offers a new possibility for the

usage of GNS as reinforcement, feasibility of the milling and the consolidation process to obtain
EP

the Cu-GNS composites with improved physical properties.


C
AC
ACCEPTED MANUSCRIPT

2. Experimental procedure

For the composites preparation, commercial Cu powder (99.9%, <75 µm) purchased from

Alfa Aesar was used. The graphene used for the composite preparation were synthesized in our

laboratory. All the chemical reactants were purchased from Merck and used without any further

PT
purification. The typical synthesis of the graphene is as follows: Firstly, the acid treated graphite

RI
was expanded using commercial microwave oven. The expanded graphite was oxidized and

consequently used for the synthesis of graphene. For the synthesis, 1 g of expanded graphite was

SC
mixed with 4 g of KMnO4. In the next step (in ice water bath) 30 ml of H2SO4 was added in slow

manner and stirred for ~20 min. The temperature was raised to 35o C and the mixture was stirred

U
until to form a thick paste. To this thick paste 100 ml of water was added under rapid stirring and
AN
the reaction temperature was raised to 85o C. The final mixture was maintained under constant
M

stirring at the same temperature for ~1 hr. The solid precipitate collected by centrifuge was

washed and dried in hot air oven. For graphene preparation in a typical reaction, 50 mg of the
D

above-mentioned oxidized product was dispersed in water (100 ml) and then the mixture was
TE

transferred through testing sieve (Advantech Manufacturing, USA, U.S. Standard Testing

Sieve, 300 µm pore size). To that dispersion, ascorbic acid solution (300 mg dissolved in 30 ml)
EP

was added slowly. The final mixture was maintained under stirring at 80o C for ~2 hr. The

resultant floating black precipitate (pure graphene) was collected by filtering, washing and
C

drying.
AC

The powder mixture of Cu-1GNS (wt. %) was placed in a ZrO2 vial (250 ml) along with 10 mm

ZrO2 balls as grinding media and the milling was performed in an ultrapure argon atmosphere

using a planetary mill Retsch PM400 with a speed of 150 rpm. The ball-to-powder ratio (BPR)
ACCEPTED MANUSCRIPT

was fixed as 10:1 and 0.5 wt.% of stearic acid (C18H36O2) was added as a process control agent

(PCA). To avoid the overheating of the vial and to maintain a stable temperature, an on/off cycle

of 10 min interval was maintained during the process. The milling time range was varied from

0.5 to 4 h. The morphology and the chemical composition of the powders obtained from the

PT
milling process were studied using scanning electron microscope (SEM, JSM-6380LV) equipped

RI
with energy dispersive X-ray spectrometer (EDXS, Oxford Instruments). The X-ray diffraction

(XRD, Bruker, D4 Endeavor) measurements were carried out by using the Cu Kα line of 40kV

SC
and 20 mA within the range of 15-80° diffraction angle.

The consolidation of the powders in a vacuum (10-2 torr) was carried out by using vacuum

U
hot-pressing technique. For this a graphite die with an inner diameter of 25.4 mm was used. The
AN
graphite die loaded with the powder was placed in a hot-pressing furnace (HP-20 Thermal

Technology Systems, USA) for the thermal treatment. The temperature was raised from room
M

temperature (RT) to 100ºC with a rate of 3 °C/min and then the heating rate was maintained as 10
D

ºC/min until to reach the consolidating temperature (600º or 700ºC). Then the die was subjected
TE

to uniaxial pressure of 30 MPa for about 30 min at the required consolidating temperature. The

consolidated pellet sample was subjected to cool down to RT with a rate of 10 ºC/min. In
EP

accordance with the ASTM E373 standard, the densities of the pellets were measured with help of

Archimedes method using double distilled water. For comparison, the theoretical density of Cu-
C

1GNS composites was calculated (8.89 g/cm3) using the rule of mixtures by considering the
AC

densities of Cu and GNS as 8.96 and 2.00 g/cm3, respectively. The microhardness of the sintered

samples was obtained using Struers microhardness tester for 3 different loads such as 0.98, 1.96

and 4.90 N applied for 10 s. In this test, for each load ten indentations were made on the samples

at room temperature. The Vickers hardness was determined by the ratio of the applied load via a
ACCEPTED MANUSCRIPT

geometrically defined indenter to the contact (projected) area of the resultant impression using

the Eq. (1):


Hv = 1854.4 (1)

PT
where ‘P’ is the applied load (Kg) and ‘d’ is the indentation diagonal length (mm). The

microstructure of the sintered samples was analyzed by using the scanning electron microscope

RI
(SEM, JSM-6380LV) technique. For measuring the electrical resistivity, rectangular bar

SC
specimens were prepared from the hot pressed disks (1 inch diameter) and the measurements

(resistance, R) were carried out by using a digital low resistance ohmmeter (DLRO, AVO Biddle

U
247002). The electrical resistivity (ρ) was calculated using the Eq. (2):
AN
×
= (2)

M

where ‘R’ is the resistance, ‘A’ is the area and ‘L’ is the length of the specimen, respectively. The

samples were etched in a solution of potassium dichromate:sulfuric acid (K2Cr2O7:H2SO4-) to


D

reveal the grain boundaries.


TE
C EP
AC
ACCEPTED MANUSCRIPT

3. Results and discussions

3.1 Characterization of the morphology and microstructure of the powders

The morphology of the graphene and Cu raw powders used for the ball milling process are shown

PT
in Figure 1(a&b) and Figure 1(c&d), respectively. The Cu powder consists of particles with

irregular morphology having the dimensions of 10-50 µm. The SEM image of the graphene used

RI
for the composites formation exhibited (Figure 1a), and disagglomerated layers of graphene (≤ 5

SC
layers) with translucent ruffled morphology. The graphene powder used in this work consists of

graphene nanosheets and the details were reported elsewhere [26].

U
AN
M
D
TE
C EP
AC

Figure 1. SEM images of (a&b) graphene nanosheets (GNS) and (c&d) Cu raw powders

used for the composites formation.

For the better understanding, the influence of the milling time on the morphology, size

and shape of powders were qualitatively investigated by using the microscopy imagery technique
ACCEPTED MANUSCRIPT

and the results are shown in Figure 2. By comparison (Fig. 1c and Fig. 2a) , we observed that the

0.5 h milled powder does not exhibited any noticeable morphological changes when compared

with the morphology of the raw Cu powder. The SEM image of the 1 h milled powder exhibited

flattening which is a typical first stage of mechanical milling (MM), that signifies the

PT
predominant cold welding [27] in the powder. The observed morphological change is a

RI
characteristic behavior in the milling of ductile powders [28]. In general, during MM the powder

particles often go through cold-welding, fracturing and re-welding processes. Also, the cold

SC
welding and fracturing processes can be predominant at any stage and they mostly depended on

the deformation characteristic of the powder precursors and their kinetics. The observed

U
appearance of small particles, along with big morphological structures of the 2 h milled powder,
AN
signifies that they are released from the coarsened particles. The further increase of milling time

(for 4 h of milling) led to the increased quantity of the small particles, which is due to the
M

predominant fracturing of the coarsened particles.


D

The observed size reduction exhibited a contradictory tendency of the Cu to form larger particles,
TE

and similar observation was reported by Yue et al. [33] in their study on bulk graphene–copper

composites. Varol et al. [29] reported that this phenomena occurred due to the lubrication effects
EP

of GNS and the shorter milling times. The addition of PCA minimized cold welding and
C

prevented agglomeration effectively. The Cu powder precursor with flakes like morphology
AC

when used for the composite formation exhibited improved functional properties previously [29].

Thus, the powders obtained from the 4 h milling are used for the consolidation at different

temperatures.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 2. SEM images of 1 wt.% GNS reinforced Cu powders formed using the (a) 0.5, (b)
M

1, (c) 2 and (d) 4 hours of milling time.


D

The XRD patterns of the GNS, pure Cu and Cu-1GNS composites sintered at 600° and 700°C are
TE

shown in Figure 3a, Figure 3b and Figure 3c, respectively. The XRD pattern of the graphene

powder exhibited peaks related to (002) and (100) reflections of graphene, which resembles a
EP

typical XRD pattern of FLG [26]. The observed broad peak corresponds to (002) reflection of

graphene was fitted and we observed that the calculated interlayer spacing is approximately
C

0.483-0.337 nm [26], which is quite close to the commercial graphite (0.336 nm) interlayer
AC

spacing. From the XRD data analysis it is observed that the graphene has 3, 4 and 24 number of

layers indicating the successful formation of FLG [26]. The appearance of few layer graphene

sheets was due to weak van der Waals forces which led to the stacking of the individual sheets

during the synthesis.


ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

Figure 3. X-ray diffraction patterns of a) GNS, b) Cu and Cu-1GNS consolidated at 600°C

and c) Cu and Cu-1GNS consolidated at 700°C.


EP

The XRD patterns of the Cu-1GNS samples, consolidated at 600 and 700°C, depicts diffraction
C

peaks at 43.2, 50.3 and 74.1°, which corresponds to (111), (200) and (220) reflections of Cu,
AC

respectively [30]. No significant differences are observed among the patterns depending on the

sintering temperature, except the lightly increased intensity of the peaks at the higher sintering

temperature, which evidences a higher crystallinity and grain growth of the particles under

sintering. The absence of graphene related diffraction peaks in Cu-1GNS samples is due to the:

(a) dispersion of low amount of graphene in the Cu matrix which is undetectable with in the
ACCEPTED MANUSCRIPT

detection limits of XRD, (b) masking of graphene related peaks by the high intense diffraction

peaks related Cu and (c) deterioration of stacking of the graphene sheets during the milling which

may result in well dispersed monolayer GNS in the Cu matrix [31]. A similar type of XRD

pattern behavior was reported on bulk graphene–copper composites [33]. Moreover, the absence

PT
of any additional diffraction peaks corresponds to oxides or carbides indicates the successful

RI
formation of the composites with noticeable chemical compatibility to produce new Cu-GNS

composite materials.

SC
3.2 Characterization of the consolidated Cu-GNS composites

U
The deformation as well as a fracture caused by the application of a load is considered as
AN
a measure of the hardness of materials. The variations of the hardness with respect to the applied

load for the Cu-GNS composites consolidated at 600 and 700oC is shown in Figure 4. The
M

measured hardness values are tabulated in Table 1 for comparison. From this study, it is evident
D

that a gradual increment of the hardness with respect to the applied load is observed and its
TE

reaches its maximum for 4.9 N of applied load.


C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 4. Dependence of hardness on applied load for Cu-GNS composites sintered at

different temperatures.
M

A clearly appreciable difference in the microstructure of the Cu-1GNS (Figure 5a) and the
D

Cu-0GNS (Figure 5b) is observed from SEM micrographs. It is evident that the microstructure of
TE

the Cu-1GNS composite depicted the agglomeration of the GNS (Figure 5a), which altered the

mechanical and electrical properties of the Cu-GNS composites [32]. The observed
EP

agglomeration of GNS acted as virtual pores, inhibiting the full densification. A similar

microstructural identification was reported by Varol et al. [29] in their study on Cu-MLG
C

composites, reporting higher MLG agglomeration than our present work, that signifies the
AC

feasibility of the usage of hot pressing for the consolidation of Cu-GNS composites. On the other

hand, the consolidated pure Cu sample just exhibited sintering pores (Figure 5b) and it is similar

with the microstructures reported by Wejrzanowski et al. [33].


ACCEPTED MANUSCRIPT

PT
RI
SC
Figure 5. SEM images of Vickers indentation, depicting the effects of GNS as reinforcement

on the microstructure of Cu-GNS composites (a) Cu-1GNS and (b) Cu-0GNS which are

U
consolidated at 600ºC. The area marked with the red circle depicts the agglomeration of the
AN
GNS in the Cu-1GNS and the area marked with white circle depicts the sintering pores in

pure Cu (Cu-0GNS). The red arrow marks shows small graphene agglomerates.
M

We also observed smaller graphene agglomerates which are located in the grain
D

boundaries (Figure 5a). For further analysis, EDS line scan analyses were conducted on the grain
TE

boundaries of the Cu-1GNS samples and the details are presented in Figure 6. The EDS analyses

clearly confirmed the presence of the graphene in the grain boundaries of the Cu-1GNS
EP

composites. Further, the BSE images from SEM for the Cu-1GNS composite are shown in

Figures 6, exhibiting the presence of the graphene agglomerates with 1- 4 µm of wide. The
C

macroscopic view of the samples as shown in Figures 6 (e&f) reveals the presence of in-
AC

homogeneously distributed graphene which is situated into the grain boundaries.


ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

Figure 6. (a) SEM micrograph and (b) EDS linescan analysis of the etched Cu-1GNS
EP

consolidated at 600°C. (c) EDS linescan analysis and (d) SEM micrograph of the Cu-1GNS

consolidated at 700°C. SEM images in back scattered mode corresponding to the


C

consolidating temperatures of (e) 600 and (f) 700°C, respectively.


AC

A comparative plot of the hardness and electrical resistivity of the consolidated Cu-GNS

composites is shown in Figure 7. It is observed that, by comparing the samples consolidated at

600°C, the addition of 1 wt.% of GNS into the copper matrix improved the electrical conductivity

around 22% when compared with the pure Cu, while maintaining their mechanical properties. In

comparison with the previously reported work on Cu-1GNS composites synthesized by hot
ACCEPTED MANUSCRIPT

pressing (800°C, 25 MPa, 30 min) [31], we achieved an increased conductivity values even with

the low temperature and slightly higher applied load used in this investigation (600°C, 30 MPa,

30 min). It is noticeable that at the higher sintering temperature, the electrical conductivity of the

Cu-1GNS composites decreased around 14%, which is attributed to the lowering of relative

PT
density by 1% (Table 1).

RI
U SC
AN
M

Figure 7. Variation of (a) hardness and (b) electrical resistivity with the loading amount of
D

GNS and for different consolidating temperatures in Cu-GNS composites consolidated by


TE

hot pressing.

The hardening of the copper matrix and increase of its electrical resistivity was reported
EP

by Dutkiewicz et al. [34] in their study of Cu-GNP (510°C and 550 MPa). The hardnesses values

obtained for the Cu-1GNS composites consolidated at 600 and 700°C in the present study (62.9
C

and 54.6 HV, respectively) are higher than the value previously reported [35] for Cu-1GNS (40
AC

HV) with hot pressing consolidating parameters as 850°C and 25 MPa for 1 hour. Thus, it

suggests that the consolidating temperature influences higher than the applied load on the

properties of Cu-GNS composites. On the other hand, a decrease in hardness value with increased

loading of multilayered graphene (MLG) was reported, due to higher agglomeration and soft

nature of MLG [29].


ACCEPTED MANUSCRIPT

Table 1. The measured hardness, relative density, IACs (%) and electrical resistivity values

of Cu-GNS composites consolidated under different temperatures.

Cu-0GNS Cu-1GNS
Parameters
600°C 700°C 600°C 700°C

PT
Hardness (HV) 63.5 48.5 62.9 54.6

RI
0.623 0.47 0.617 0.55
Hardness (GPa)
(± 0.07) (± 0.03) (± 0.02) (± 0.01)

SC
Relative density
83.5 88.4 92.6 91.6
(%)

IACs (%) 77
U 99 94 81
AN
2.22 1.73 1.83 2.13
R (Ω.m X 10-8)
M

(± 0.04) (± 0.02) (± 0.03) (± 0.03)


D

The revealed microstructures after etching of the Cu-GNS composites are shown in the
TE

Figure 8. The consolidating temperatures considered in the current study are under 750°C, in

which the grain growth of Cu is slow and limited [36]. The Cu-1GNS composites consolidated at
EP

600 and 700 °C exhibited a similar grain size, whereas, the pure Cu (Cu-0GNS) sample

consolidated at 600 °C exhibited definite grain boundary and it is absent in the pure Cu (Cu-
C

0GNS) consolidated at 700 °C, which is related to the observed decrease in hardness of the
AC

consolidated pure Cu sample (Cu-0GNS) with the increase in temperature. For consolidation at

700°C the observed increase in the hardness for the composite samples (Cu-1GNS) when

compared with Cu-0GNS is due to the ability of GNS to delimit the grains effectively [37]. This

signifies that GNS interface in Cu matrix has a noteworthy influence on hardness of the powders

when consolidated.
ACCEPTED MANUSCRIPT

The relative density of Cu-1GNS composites measured by the Archimedes method (Table

1) exhibited an increment of ∼10 and ∼3% when compared with the Cu-0GNS (pure Cu) for the

consolidating temperatures of 600 and 700 °C, respectively. This supports the observed in the

increase in the quantity of pores (Figure 8) in the Cu-1GNS composites. It is observed that the

PT
relative density of the Cu-1GNS is increased when compared to Cu-0GNS for the corresponding

RI
consolidating temperatures, being contrary to observations reported by Varol et al. [29] but is in

well agreement with the Wei et al. [31]. The mechanical properties are strongly influenced by the

SC
porosity and according to the Gibson and Ashby who proposed an equation which relates these

variables, stated that the mechanical properties will decreases when the porosity level increases

U
[38]. For high temperature industrial applications [39], the Cu-GNS composites should have
AN
IACs value over 50. The Cu-1GNS from the present work exhibited higher value of IACs when
M

compared to the previously reported IACs value of the 1 wt.% MLG reinforced Cu composite

[29].
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 8. SEM images of Cu-GNS composites consolidated at different temperatures (a) Cu-1GNS,
M

600°C ; (b) Cu-1GNS, 700°C ; (c) Cu-0GNS, 600 °C and (d) Cu-0GNS, 700°C.
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

4. Conclusions

The present work demonstrated that the graphene loading into the Cu matrix clearly influenced

the mechanical and electrical conductivity by varying the sintering temperature. For instance, the

graphene loaded Cu composites prepared by sintering at 600 oC exhibited a nominal change in

PT
the mechanical property with a significant increase in the electrical conductivity when compared

RI
to the pure Cu sample. Relevantly, composites obtained by sintering at 700oC exhibited

improved mechanical property with a slight decrease in the conductivity. Also, the observed

SC
decrease in the in hardness of the consolidated pure Cu sample (Cu-0GNS) with the increase in

temperature was due to the grain boundary variation with the temperature. Due to the lubrication

U
effect of GNS, a predominance of the fracture in the surface of the flattened powders which is
AN
against the tendency of copper to cold-welding was observed in the powders mechanically milled

for 4h. The composite processed at 700°C exhibited lower hardness values than composite
M

processed at 600°C due to the increased grain size. By adding 1 wt.% of GNS to the copper, the
D

electrical conductivity was improved around 22% by comparing with the pure Cu consolidated at
TE

600 oC. Whereas, 18% of reduction in the electrical conductivity was observed for 700 oC

consolidating temperature which was due to the both grain morphology and grain size produced
EP

by adding 1wt.% GNS to the copper matrix. The relative density of Cu-1GNS composites

exhibited an increment of ~10% and ~3% when compared with the Cu-0GNS (pure Cu) for the
C

corresponding consolidated temperatures. In comparison with the previously reported works on


AC

Cu-GNS composites, we achieved increased conductivity values of around 22% even with the

lower temperature and higher applied load in the consolidation (600°C, 30 MPa, 30 min).

Concluding that reducing the sintering temperature and increasing the applied pressure in the hot

pressing process are suitable for the realization of Cu-GNS composites with enhanced physical

and mechanical properties.


ACCEPTED MANUSCRIPT

5. Acknowledgements

The authors gratefully acknowledges the FONDEQUIP/EQM N°130103. C. Salvo CONICYT for

financing his PhD studies (grant CONICYT-PCHA/Doctorado Nacional/2016-21161251).

PT
6. References

RI
1. Rajkovic, V., D. Bozic, and M.T. Jovanovic, Properties of copper composites
strengthened by nano-and micro-sized Al2O3 particles. International Journal of

SC
Materials Research, 2010. 101(3): p. 334-339.
2. Hong, E., et al., Tribological properties of copper alloy-based composites reinforced
with tungsten carbide particles. Wear, 2011. 270(9-10): p. 591-597.
3. Shehata, F., et al., Preparation and properties of Al2O3 nanoparticle reinforced copper

U
matrix composites by in situ processing. Materials & Design, 2009. 30(7): p. 2756-
2762.
AN
4. Moghadam, A.D., et al., Functional metal matrix composites: self-lubricating, self-
healing, and nanocomposites-an outlook. Jom, 2014. 66(6): p. 872-881.
5. Ferguson, J., et al., On the strength and strain to failure in particle-reinforced
M

magnesium metal-matrix nanocomposites (Mg MMNCs). Materials Science and


Engineering: A, 2012. 558: p. 193-204.
6. Moghadam, A.D., et al., Mechanical and tribological properties of self-lubricating metal
D

matrix nanocomposites reinforced by carbon nanotubes (CNTs) and graphene–a


review. Composites Part B: Engineering, 2015. 77: p. 402-420.
7. Lu, L., et al., Ultrahigh strength and high electrical conductivity in copper. Science,
TE

2004. 304(5669): p. 422-426.


8. Duarte, I. and J.M. Ferreira, Composite and nanocomposite metal foams. Materials,
2016. 9(2): p. 79.
EP

9. Tjong, S.C., Recent progress in the development and properties of novel metal matrix
nanocomposites reinforced with carbon nanotubes and graphene nanosheets.
Materials Science and Engineering: R: Reports, 2013. 74(10): p. 281-350.
10. Geim, A.K. and K.S. Novoselov, The rise of graphene. Nature materials, 2007. 6(3): p.
C

183.
AC

11. Huang, X., et al., Graphene-based composites. Chemical Society Reviews, 2012. 41(2):
p. 666-686.
12. Novoselov, K.S., et al., Electric field effect in atomically thin carbon films. science,
2004. 306(5696): p. 666-669.
13. Ponraj, N.V., et al., Graphene nanosheet as reinforcement agent in copper matrix
composite by using powder metallurgy method. Surfaces and Interfaces, 2017. 6: p.
190-196.
14. Nieto, A., D. Lahiri, and A. Agarwal, Synthesis and properties of bulk graphene
nanoplatelets consolidated by spark plasma sintering. Carbon, 2012. 50(11): p. 4068-
4077.
ACCEPTED MANUSCRIPT

15. Lee, C., et al., Measurement of the elastic properties and intrinsic strength of monolayer
graphene. science, 2008. 321(5887): p. 385-388.
16. Frank, I., et al., Mechanical properties of suspended graphene sheets. Journal of
Vacuum Science & Technology B: Microelectronics and Nanometer Structures
Processing, Measurement, and Phenomena, 2007. 25(6): p. 2558-2561.
17. Li, J.-f., et al., Sliding wear behavior of copper-based composites reinforced with
graphene nanosheets and graphite. Transactions of Nonferrous Metals Society of

PT
China, 2015. 25(10): p. 3354-3362.
18. Koltsova, T.S., et al., New hybrid copper composite materials based on carbon
nanostructures. Journal of Materials Science and Engineering B, 2012. 2(4): p. 240-

RI
246.
19. Pavithra, C.L., et al., A new electrochemical approach for the synthesis of copper-
graphene nanocomposite foils with high hardness. Scientific reports, 2014. 4: p. 4049.

SC
20. Wang, L., et al., Graphene-copper composite with micro-layered grains and ultrahigh
strength. Scientific reports, 2017. 7: p. 41896.
21. Gao, X., et al., Mechanical properties and thermal conductivity of graphene reinforced
copper matrix composites. Powder Technology, 2016. 301: p. 601-607.

U
22. Jiang, R., et al., Copper–graphene bulk composites with homogeneous graphene
dispersion and enhanced mechanical properties. Materials Science and Engineering: A,
AN
2016. 654: p. 124-130.
23. Kumar, H.P. and M.A. Xavior, Graphene reinforced metal matrix composite (GRMMC):
a review. Procedia Engineering, 2014. 97: p. 1033-1040.
M

24. Sharifi, E.M., F. Karimzadeh, and M. Enayati, Fabrication and evaluation of mechanical
and tribological properties of boron carbide reinforced aluminum matrix
nanocomposites. Materials & Design, 2011. 32(6): p. 3263-3271.
D

25. Li, W., et al., Conductive enhancement of copper/graphene composites based on high-
quality graphene. RSC Advances, 2015. 5(98): p. 80428-80433.
TE

26. Udayabhaskar, R., et al., Spectroscopic investigation on graphene-copper


nanocomposites with strong UV emission and high catalytic activity. Carbon, 2017.
124: p. 256-262.
27. Intrater, J., Mechanical alloying and milling, c. Suryanarayana. 2007.
EP

28. Benjamin, J.S. Fundamentals of mechanical alloying. in Materials Science Forum. 1992.
Trans Tech Publ.
29. Varol, T. and A. Canakci, Microstructure, electrical conductivity and hardness of
C

multilayer graphene/copper nanocomposites synthesized by flake powder metallurgy.


Metals and materials international, 2015. 21(4): p. 704-712.
AC

30. File, P.D., International centre for diffraction data. Swarthmore, Pa, 2000. 19081.
31. Wei, B.W., et al. Synthesis and Physical Properties of Graphene Nanosheets Reinforced
Copper Composites. in Advanced Materials Research. 2014. Trans Tech Publ.
32. Wang, C., et al., Fabrication, interfacial characteristics and strengthening mechanisms
of ZrB2 microparticles reinforced Cu composites prepared by hot-pressed sintering.
Journal of Alloys and Compounds, 2018. 748: p. 546-552.
33. Wejrzanowski, T., et al., Thermal conductivity of metal-graphene composites.
Materials & design, 2016. 99: p. 163-173.
ACCEPTED MANUSCRIPT

34. Dutkiewicz, J., et al., Microstructure and properties of bulk copper matrix composites
strengthened with various kinds of graphene nanoplatelets. Materials Science and
Engineering: A, 2015. 628: p. 124-134.
35. Yue, H., et al., Effect of ball-milling and graphene contents on the mechanical
properties and fracture mechanisms of graphene nanosheets reinforced copper matrix
composites. Journal of Alloys and Compounds, 2017. 691: p. 755-762.
36. Zhang, Z.-H., et al., Ultrafine-grained copper prepared by spark plasma sintering

PT
process. Materials Science and Engineering: A, 2008. 476(1-2): p. 201-205.
37. Ponraj, N.V., et al., Effect of milling on dispersion of graphene nanosheet reinforcement
in different morphology copper powder matrix. Surfaces and Interfaces, 2017. 9: p.

RI
260-265.
38. Gibson, L.J. and M.F. Ashby, Cellular solids: structure and properties. 1999: Cambridge
university press.

SC
39. Grant, N., A. Lee, and M. Lou, High conductivity copper and aluminum alloys. American
Institute of Mining, Metallurgical, and Petroleum Engineers, Inc., New York, 1984: p.
103.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Highlights
1. Loading of graphene led to the 22% of rise in electrical conductivity of the Cu.
2. Graphene induced lubrication led to the predominance of the fracture.
3. Relative density exhibited an increment with graphene loading.
4. Low sintering temperature and high applied load are optimal for consolidation.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like