You are on page 1of 9

Arabian Journal for Science and Engineering

https://doi.org/10.1007/s13369-021-05367-3

RESEARCH ARTICLE-MECHANICAL ENGINEERING

A Micromechanics‑Based Hierarchical Analysis of Thermal


Conductivity of Metallic Nanocomposites with Agglomerated Ceramic
Nanoparticles
M. K. Hassanzadeh‑Aghdam1 · R. Ansari2

Received: 27 July 2020 / Accepted: 13 January 2021


© King Fahd University of Petroleum & Minerals 2021

Abstract
The influence of agglomeration of SiC nanoparticles on the bulk thermal conducting behavior of aluminum (Al)-based
nanocomposites is examined using a micromechanics-based hierarchical technique. The interfacial thermal resistance (ITR)
between the ceramic nano-scale particles and the metallic matrix is included in the analysis. The predictions of the microme-
chanical model considering the agglomeration and ITR are in very good agreement with the available experimental results.
The agglomeration of ceramic nanoparticles greatly reduces the thermal conducting coefficient of the Al-based nanocom-
posites. When the nanoparticle volume fraction is 10%, the thermal conductivity decreases from 132 to 115 W/mK with the
formation of nanoparticle agglomeration. The uniform distribution of nanoparticles and the elimination of ITR can lead to a
substantial enhancement in the nanocomposite thermal conductivity. When the volume fraction is 10%, the thermal conduc-
tivity increases from 115 to 188 W/mK by uniform dispersing the nanoparticles and removing the SiC/Al ITR. Moreover,
the effects of amount, and diameter of nano-scale particles as well as the constituent material properties on the bulk thermal
conductivity of the SiC nanoparticle-filled Al nanocomposites are examined. When the SiC diameter increases from 35 nm
to 3.5 µm, the thermal conductivity of the Al-based composite increases from 132 to 173 W/mK.

Keywords Aluminum-based nanocomposite · Thermal conductivity · Agglomeration · Ceramic nanoparticle · Hierarchical


modeling

1 Introduction superior high-temperature mechanical properties and low


cost [7–9].
Aluminum (Al) and its alloys have attracted much attention It has been confirmed that the nanoscale ceramic particles
for the production of the structural materials in the aero- are more favorable than the microscale ceramic particles to
space, automotive and electronic industries because of good increase the mechanical characteristics of the metallic com-
mechanical, physical and chemical properties [1–3]. In order posite materials [10, 11]. For example, Keneshloo et al. [11]
to improve the mechanical properties such as the stiffness explored that the Al-4% Cu matrix nanocomposites filled
and strength, ceramic particles are added in the Al-based with SiC nanoparticles (30 nm) indicate higher compres-
materials forming the Al-based composites [4–6]. The SiC sive strength, higher hardness, and better wear properties
particles are commonly offered as the reinforcement phase as compared to the Al-4% Cu matrix composites filled with
for the metallic composites due to the high wear resistance, SiC microparticles (15 µm). The ceramic nanoparticle-rein-
forced metallic nanocomposites can be considered as the
cutting-edge materials for advanced structural applications
* M. K. Hassanzadeh‑Aghdam [12]. Besides, they have potential applications in the thermal
mk.hassanzadeh@gmail.com protection system, especially for the thermal shock environ-
1 ment [13–15].
Department of Engineering Science, Faculty of Technology
and Engineering, East of Guilan, University of Guilan, Many research works were conducted to evaluate the
Rudsar‑Vajargah, Iran mechanical properties of the ceramic nanoparticle-filled
2
Faculty of Mechanical Engineering, University of Guilan, metallic nanocomposites. Kollo et al. [16] observed that
Rasht, Iran increasing the amount of nanoscale SiC particles from 1 to

13
Vol.:(0123456789)
Arabian Journal for Science and Engineering

10 vol.% caused the tensile strength of the Al-based nano- composites. Hasselman et al. [26] tested the thermal con-
composites to improve from 205 to 420 MPa. Murthy et al. ductivity of a 40 vol% SiC/Al composite as a function of
[17] fabricated some specimens of the Al-based nanocom- mean particle size ranging from 700 nm to 28 µm. The
posites reinforced with 0.5, 1, 1.5 and 2 wt% SiC particles reduction in thermal conductance property was observed
with a mean size of 50 nm and 150 nm. The results revealed by the decrease in SiC size. This behavior was attributed
that the ultimate tensile strength and hardness of the Al- to the interfacial thermal barrier at the SiC/Al interface.
based nanocomposites enhance with increasing the SiC Molina et al. [28] produced the Al-based composites rein-
weight fraction [17]. Mazaheri and Shabani [18] evaluated forced with mixtures of diamond and SiC particles. It was
the mechanical properties of the Al-Si alloys reinforced with confirmed that replacing the SiC gradually by the diamond
SiC nanoparticles. It was shown that adding the SiC nano- filler leads to a steady enhancement of the thermal con-
particles enhances the hardness, stiffness, yield strength and ductivity from 220 to 580 W/mK. El-Kady and Fathy [29]
tensile strength of the metallic nanocomposites. The maxi- investigated the effects of size and weight fraction of SiC
mum values of the yield strength and tensile strength were particles on the thermal conductivities of the Al-based
observed by adding 3.5 vol% SiC nanoparticles [18]. El- composite materials. Composite specimens with 5 and
Daly et al. [19] obtained the quantitative information about 10 wt% SiC nanoparticles were produced by the powder
the influence of nanoscale SiC particles on the elastic prop- metallurgy technique followed by the hot extrusion. It was
erties such as the bulk modulus, shear modulus, Young’s observed that the thermal conductivity of the Al-based
modulus and Poisson’s ratio of Al-based nanocomposites nanocomposites reduced with increasing the amount of
using the pulse echo overlap method. The results revealed SiC nanoparticles and improved with increasing SiC par-
that Young’s modulus of 10 vol% SiC nanoparticle-filled ticle size [29]. Generally, the thermal conductivity of par-
Al nanocomposites is 97.1 GPa, which is much higher than ticulate metal matrix nanocomposites (PMMNCs) is as a
that of the pure Al material (72.6 GPa) [19]. Abdullahi and function of volume fraction, size and dispersion type of
Al-Aqeeli [20] investigated the effect of SiC addition on nanoparticles, constituent properties, and especially ITR
the final morphology and microstructure of the Al-based between the nanoparticle and metal matrix [30–33]. It is
nanocomposite. Boostani et al. [21] produced the Al-based difficult to completely evaluate the thermal properties of
nanocomposites reinforced with nano-SiC particles. Based PMMNCs by the experimental methods due to their com-
on the results, 45% and 84% improvement have been found plexity and time-consuming. In this frame, the theoreti-
for the yield strength and tensile ductility. The bulk proper- cal techniques such as the micromechanical models are
ties of metallic nanocomposites are controlled by the size required for predicting the thermal conductivities of the
and volume fraction of the nanoparticle reinforcement as ceramic nanoparticle-reinforced metallic nanocomposites.
well as the constituent material properties [18–20]. The high A literature survey shows that the relation between
mechanical properties can be achieved when the ceramic the microstructural features and the macroscopic ther-
nanoparticles are dispersed uniformly in the metallic nano- mal conductivity of the SiC nanoparticle-reinforced Al-
composites. Although a number of research works exist in based nanocomposites has not been suitably presented.
developing the fabrication techniques, the homogeneous dis- Thus, it is needed to develop a micromechanical method
tribution of nano-scale particles is still a problem. Several for the quantitative characterization of the content, size
reasons such as the great surface-to-volume ratio, attractive and agglomeration of ceramic nanoparticles as well as the
van der Waals (vdW) interactions, and poor wettability of ITR. In this work, the thermal conducting response of the
fine ceramic fillers by the molten Al lead to the formation SiC nanoparticle-reinforced Al-based nanocomposites is
of an agglomerated state and inhomogeneous distribution of studied using a multi-step micromechanical approach with
the nanoparticles [21–24]. high accuracy and efficiency. One of the main advantages
The thermal properties of the metallic composites have of the proposed method is that the micromechanics model
a significant effect on the structure durability. Miniaturiza- is totally analytical. This micromechanical model can
tion of the electronic constituents has led to a growth of include the critical microstructural features such as the
power density in electronic devices which need to a great amount, size and agglomeration of nanoparticles and the
thermal conducting coefficient of the thermal management nanoparticle/metal ITR. It is shown that the uniform dis-
materials to guarantee the heat dissipation [25–28]. The persion of nanoparticles and the elimination of ITR may
primary concern in the thermal management applications lead to a significant improvement in the thermal conduc-
is identified to be the high thermal conductivity. Xu et al. tivity. Moreover, increasing the size of SiC nanoparticles
[8] measured the thermal conductivity of 10, 20, 30 and 40 improves the thermal conducting performance of the Al-
vol% SiC/Al composites. It was found that the interfacial based nanocomposites. The model predictions are in good
thermal resistance (ITR) between the SiC and Al matrix agreement with the experimental results.
may decrease the thermal conductivity of the metallic

13
Arabian Journal for Science and Engineering

2 Micromechanical Modeling of PMMNCs Vinclusion V inclusion


𝜉= ,𝜁 = r 0 ≤ 𝜉, 𝜁 ≤ 1 (2)
V Vr
The agglomeration of ceramic nanoparticles is formed
during the fabrication of the metallic nanocomposites where Vrinclusion denotes the volume of inclusions in the
[21–24, 34]. An image of the Al-based nanocomposite composite RVE. The volume fraction of spherical inclu-
which contains the agglomeration of aluminum oxide sions regarding the whole volume of composite RVE (V )
nanoparticles has been shown in Ref. [34]. In order to is signified by 𝜉 , and the volume ratio of the nanoparticles
investigate the influence of nanoparticle agglomeration into the spherical inclusions over the whole nanoparticles;
on the thermal conductivity of PMMNCs, a multi-step i.e., Vr , is indicated by 𝜁 . The ceramic nanoparticles are
homogenization technique is developed. This process uniformly dispersed in the composite material when 𝜉 = 1.
was formerly employed to obtain the elastic modulus of However, the degree of nanoparticle agglomeration becomes
the polymer nanocomposites containing carbon nanotube more severe by the decrease of 𝜉 . Also, 𝜁 = 1 indicates that
agglomeration [35, 36]. The spatial dispersion of ceramic all the ceramic nanoparticles agglomerate in the spherical
nanoparticles in the metal matrix is non-uniform such that inclusions. A uniform dispersion of nanoparticles in the
some local areas have a greater concentration of nanopar- composite materials exists when 𝜉 = 𝜁 . The volume frac-
ticles than the average volume fraction in the material, as tion of the ceramic nanoparticles in the composite material
shown in Fig. 1. is expressed as
A number of nanoparticles are uniformly dispersed in
Vr
the metallic matrix formed the effective matrix phase, and vr = (3)
the remaining nanoparticles are observed in agglomerated V
state in spherical inclusions. In the effective matrix phase On the basis of Eqs. (1–3), it can be expressed [35, 36]
and inclusion, the nanoparticles are uniformly dispersed
in the metal matrix. The thermal properties of the spheri- Vrinclusion 𝜁 Vrm (1 − 𝜁)
vinclusion = = vr , vm = = v
cal inclusions containing agglomerated nanoparticles are r Vinclusion 𝜉 r V − Vinclusion (1 − 𝜉) r
different from the surrounding material. In the proposed (4)
method, the total volume of ceramic nanoparticles Vr in where Vrinclusion and vm
r are the volume fractions of ceramic
the representative volume element (RVE) of the composite nanoparticles in the inclusion and in the metallic matrix
system is divided into two parts, as follows [35, 36] outer the inclusions, respectively.
The PMMNC is considered as a composite system con-
Vr = Vrinclusion + Vrm (1)
sisting of spherical inclusions embedded in the effective
where Vrinclusion and Vrm specify the volumes of nanoparti- matrix phase. Both effective matrix and inclusion phases
cles dispersed in the spherical inclusions and in the metallic contain ceramic nanoparticles with a uniform dispersion.
matrix outside the spherical inclusions (i.e., effective matrix The effective thermal conducting coefficients of the spheri-
phase), respectively. Two parameters 𝜉 and 𝜁 are introduced cal inclusion, the equivalent matrix, and then the PMMNCs
to reflect the degree of nanoparticle agglomeration, as must be computed.
follows
2.1 Mori–Tanaka Model

To predict the thermal conductivities of the composite


materials, the Mori–Tanaka (M-T) micromechanics model
Spherical inclusion
is employed [37–39]. In the case of steady-state heat conduc-
tion issues, the M-T method considers a single ellipsoidal
Nanoparticle
heterogeneity inserted into an infinite homogeneous matrix
domain under a constant far-field heat flux [40]. The heat
flow in the composite system is expressed as follows
Nanoparticle

̂
q = −K∇T (5)
Al matrix
where 𝐊̂ is the second-rank thermal conductivity tensor, 𝐪
and ∇T are the average of heat flux vector and temperature
Al matrix gradient, respectively. In the M-T model, it is assumed that
the resultant temperature gradient in each heterogeneity is
Fig. 1  Model of a PMMNC with ceramic nanoparticle agglomeration constant [37–40]. The thermal conductivity tensor of the

13
Arabian Journal for Science and Engineering

composite materials is explicitly determined using the fol- of metallic material and the ceramic nanoparticles outside
lowing relation the spherical inclusion. The effect of ITR is considered in
}−1 the micromechanical modeling. The bulk thermal conductiv-
K = K M . I + cR (S − I).(A − S)−1 . I + cR S.(A − S)−1 ity of the PMMNCs is computed by the M-T method using
{ }{

(6) the material properties of the effective matrix phase and the
in which spherical inclusion.
)−1
A = K M − K R .K M (7)
(
3 Results and Discussion
where 𝐊M and 𝐊R refer to the second-rank thermal con-
ductivity tensors for the metallic matrix and reinforcement, The validity of the micromechanics-based hierarchical
respectively, cR is the reinforcement volume fraction, and method is verified. To this end, the thermal conductivities
𝐈 denotes the second-rank identity matrix. Also, 𝐒 denotes of the SiC nanoparticle-reinforced Al-based nanocomposites
the second-rank Eshelby tensor which its components for predicted by the micromechanical model are compared with
various types of reinforcement can be readily found in Refs. the experimental results [29]. Thermal conductivity of this
[40, 41]. PMMNC has been evaluated experimentally by El-Kady and
Fathy [29]. In this verification, the thermal conductivities
2.2 Effective Medium Model of the Al matrix and SiC particles are 178 and 300 W/mK,
respectively. In the micromechanical modeling, the value of
In this section, the thermal conductivity of the nanoparti- ITR is 6.85 × 10–9 ­m2K/W. Figure 2 exhibits the comparison
cle-reinforced composites is analyzed. The ITR between between two sets of results for the nanoscale SiC particle-
the nanoparticles and matrix significantly affects the heat filled Al nanocomposites.
transfer performance of the composites containing nano- The variation of the effective thermal conductivity with
scale particles [30–33]. The ITR has been categorized as the weight fraction of nanoparticle is plotted. The nanopar-
a heat flow barrier associated with a weak contact at the ticle weight fraction is converted into the volume fraction in
interface, diversities in phonon spectra on the basis of the the micromechanical simulation by the densities of SiC and
atomic arrangements and densities of the phases [42]. Al matrix. The influences of important features, i.e., the SiC
The effective medium approach [30] can predict the nanoparticle agglomeration and the ITR on the thermal con-
thermal conductivities of the particulate composites. The ductivity of the PMMNCs are investigated in Fig. 2. There is
modified form of this method considering the ITR is used, a large difference between the experimental measurements
as follows [29] and the estimations of the micromechanical method
without the ITR and the nanoparticle agglomeration as
shown in Fig. 2a. Furthermore, Fig. 2b reveals that the pre-
[ ]
NC M
K NP (1 + 2𝛼) + 2K M + 2cNP K NP (1 − 2𝛼) − K M
K =K [
K NP (1 + 2𝛼) + 2K M − cNP K NP (1 − 2𝛼) − K M
] dictions of the thermal conductivity with agglomeration in
(8) the absence of the ITR are greater than the experiments [29].
The dependency of the thermal conductivities of PMMNCs
where K NP is the nanoparticle thermal conductivity, and 𝛼 is
on the nanoparticle dispersion type is observed from the
a dimensionless parameter defined as
comparison of the results displayed in Fig. 2 a, b. The ther-
RK K O mal conducting coefficient of the Al-based nanocomposites
𝛼= (9) reinforced with the uniformly dispersed nanoparticles is
(d∕2)
greater than that of Al-based nanocomposites containing
where d refers to the diameter of the nanoparticle, and RK the nanoparticle agglomeration. The uniform dispersion of
is the nanoparticle/matrix ITR. The thermal property of ceramic nanoparticles helps to make the better heat dissipa-
the interfacial region is concentrated on a surface of zero tion from the structures made of PMMNCs. Considering
thickness. only the ITR is not adequate factor to have a good prediction
A three-step modeling process is performed to calculate of the thermal conductivity of the Al-based nanocomposites
the thermal conductivity of the ceramic nanoparticle-rein- as compared to the experimental data [29] as indicated in
forced metallic nanocomposites. In the first step, the thermal Fig. 2c. The comparison between the results of the microme-
conducting coefficient of the spherical inclusion consisted of chanical method presented in Fig. 2 a, c clarifies that the ITR
the ceramic nanoparticles and metallic matrix is computed significantly affects the thermal properties of the PMMNCs.
using the effective medium model. In the second step, a sim- The thermal conductivities of the Al-based nanocomposites
ilar procedure is made to calculate the thermal conducting without the ITR are very greater than those of the Al-based
coefficient of the effective matrix phase consisted of the rest nanocomposites with the ITR. Consequently, eliminating

13
Arabian Journal for Science and Engineering

(a) (b)

Thermal conductivity (W/mK) 180 180

Thermal conductivity (W/mK)


Experiment [29] Experiment [29]
155 155
Present method, Without agglomeration, Present method, With agglomeration,
Without ITR Without ITR
130 130

105 105

80 80
0 3 6 9 12 0 3 6 9 12
SiC nanoparticle amount (wt.%) SiC nanoparticle amount (wt.%)

(c) Experiment [29] (d) Experiment [29]

180 Present method, Without agglomeration, 180


Thermal conductivity (W/mK)

Thermal conductivity (W/mK)


With ITR Present method, With agglomeration,
With ITR
155 155

130 130

105 105

80 80
0 3 6 9 12 0 3 6 9 12
SiC nanoparticle amount (wt.%) SiC nanoparticle amount (wt.%)

Fig. 2  Comparison between the model predictions and experiment without ITR, c without agglomeration, with ITR, and d with agglom-
[29] of Al-based nanocomposites reinforced with nanoscale SiC par- eration, with ITR
ticles; a without agglomeration, without ITR, b with agglomeration,

the ITR can cause a substantial enhancement in the thermal


conductivity. It is observed from Fig. 2d that both agglom-
eration of SiC nanoparticles and the ITR must be included 180
Thermal conductivity (W/mK)

in the micromechanical modeling to obtain more accurate


estimations as compared to the experimental measurements 155
[29]. The value of 𝜉 is 0.3, and 𝜁 is assumed to be a volume
fraction-dependent parameter, as follows
130
𝜁 = 0.65 + 2.5cNP (10)
Experiment [29]
The increase in SiC nanoparticle content may improve 105
the PMMNC thermal conductivity by eliminating both the Present method
nanoparticle agglomeration and the SiC/Al ITR.
80
Another comparison is made between the results of the 0 3 6 9 12
micromechanical model and experimental measurements SiC particle amount (wt.%)
[29] to further show the validity of the model. The compari-
son is shown in Fig. 3. The experimental tests were given
Fig. 3  Comparison between the model predictions and experimental
results [29] of Al-based composites reinforced with SiC particles

13
Arabian Journal for Science and Engineering

for two weight fractions of SiC particles, including 5 and 1.35


10%. The variation of the thermal conductivity of the SiC/
Al composites versus the particle weight fraction is plot- FEM [43]

Normalized thermal conductivity


1.28
ted. The effects of important microstructural features are
Present method
included in the micromechanical modeling. It is found that
the predictions of the model are in good agreement with the 1.21
experimental results [29].
Figure 4 indicates the comparison between the model pre- 1.14
dictions and experimental data [8] of the thermal conductiv-
ity of the SiC/Al composites. The thermal conductivity of 1.07
the Al matrix is 174.57 W/mK. Also, in the micromechani-
cal modeling, the value of ITR is 6.85 × 10–9 ­m2K/W. The
figure shows the variation of the thermal conductivity with 1
0 2 4 6 8 10
the SiC volume fraction. Significant contribution of the ITR
Nanoparticle amount (vol.%)
between the SiC particle and the Al matrix to the thermal
conducting behavior of SiC/Al composites can be seen in
Fig. 5  Comparison between the results of the present model and the
Fig. 4. It is observed that with considering the ITR in the
FEM [43] for the thermal conductivity of the particulate nanocom-
micromechanical modeling, two sets of results are in good posites
agreement validating the theoretical method utilized in this
study.
A comparison between the results of the present analyti- Figure 6 depicts the influence of nanoparticle/metal ITR
cal micromechanics-based model and finite element method on the PMMNC thermal conductivity. The value of Rk is
(FEM) [43] is presented. Figure 5 shows the normalized chosen to be 0, 0.1 × 10–9, 1 × 10–9 and 6.85 × 10–9 ­m2K/W.
thermal conductivity of the particulate nanocomposites A homogenous dispersion of the ceramic nanoparticles into
versus the nanoparticle volume fraction. It is noticed that the Al matrix is considered. The notable sensitivity of the
the thermal conductivity of the nanocomposites is normal- thermal properties to the ITR can be found from this figure.
ized with respect to the matrix thermal conductivity. A good The maximum value of the thermal conductivity is deter-
agreement is observed between the predictions of the model mined when Rk=0. It is illustrated that the thermal conduc-
and the FEM results [43]. tivity decreases by increasing the ITR. For example, when
Parametric studies are conducted to investigate the ITR, the SiC nanoparticle weight fraction is 5%, in comparison
degree of agglomeration, size of nanoparticles, and constitu- with the case of Rk=1 × 10–9 ­m2K/W, the thermal conducting
ent material properties which mostly affect the homogenized coefficient of the PMMNC without ITR enhances by 16.6%
thermal properties of the SiC nanoparticle-reinforced Al- from 157 to 183 W/mK. Therefore, the reduction in the ITR
based nanocomposites. is necessary to obtain a higher thermal conductivity. When

230
Thermal conductivity (W/mK)

205 180
Thermal conductivity (W/mK)

180
155

155
130
130 Rk=6.85e-9 m^2K/W
Experiment [8]
Rk=1e-9 m^2K/W
Present method, With ITR 105 Rk=0.1e-9 m^2K/W
105
Present method, Without ITR Rk=0

80 80
0 8 16 24 32 40 48 0 3 6 9 12
SiC particle amount (vol.%) SiC nanoparticle amount (wt.%)

Fig. 4  Comparison between the model predictions and experiment [8] Fig. 6  Influence of ITR on the thermal conductivity of Al-based
of Al-based composites reinforced with SiC particles nanocomposites reinforced with nanoscale SiC particles

13
Arabian Journal for Science and Engineering

Rk = 0, the increase in ceramic nanoparticle content leads to


an enhancement of the PMMNC thermal conductivity.
180
Inherent characteristics of the nanoscale ceramic parti-

Thermal conductivity (W/mK)


cles and other factors induced during the production of the
PMMNCs make the SiC nanoparticles easy to agglomerate 155
in real engineering situations. The thermal conducting coef- d=35 nm
ficient of the Al-based nanocomposites as a function of SiC
130 d=350 nm
nanoparticle weight fraction for different values of 𝜁 , includ-
ing 0.5, 0.65, and 0.8 is indicated in Fig. 7. It is found an d=3.5 µm

increased thermal conductivity for the PMMNCs owing to 105 d=5 µm


the uniform dispersion of SiC nanoparticles as 𝜉 = 𝜁 = 0.5.
d=10 µm
The homogenized thermal conductivities of the PMMNCs 80
have lower values as the value of 𝜁 increases. In other words, 0 3 6 9 12
when ceramic nanoparticles form agglomeration regions SiC nanoparticle amount (wt.%)
( 𝜉 ≠ 𝜁 ) in the PMMNCs, degradation of the thermal con-
ductivity is observed against the PMMNCs containing Fig. 8  Influence of diameter of SiC particles on the thermal conduc-
well-dispersed nanoparticles. Therefore, fabrication of the tivity of Al-based composites
Al-based nanocomposites with a homogenous nanoparticle
dispersion is expected to be an efficient way to attain a more
level of thermal conductivity. This can make the heat dissi- To examine the effect of the metal matrix thermal con-
pation from structures faster which is favorable in numerous ductivity on the effective PMMNC thermal conducting
engineering applications, especially in electronic equipment. coefficient, another micromechanical analysis is performed.
The thermal conductivity of Al-based composites versus Figure 9 displays the variation of PMMNC thermal conduct-
the particle content for different values of SiC diameter is ing coefficient with the matrix thermal conductivity in the
presented in Fig. 8. For this micromechanical analysis, five range of 150–250 W/mK. The predictions have been illus-
different values, including 35 nm, 350 nm, 3.5 µm, 5 µm trated in the presence and the absence of ITR. Generally,
and 10 µm, are selected. As the ceramic particle diameter the PMMNC thermal conductivity displays a linear increase
increases, the thermal conductivity of the Al-based compos- trend with increasing the matrix thermal conductivity. As
ites can be increased. This is due to the fact that the increase expected, the thermal conductivities of Al-based nanocom-
in the particle diameter decreases the influence of thermal posites with the ITR are lower than those of the Al-based
resistance in interfacial region on the composite thermal nanocomposites without ITR.
conductivity. A threshold value of particle diameter exists A sensitivity analysis is directed to explore the role of
after which further increase in diameter does not contribute nanoscale particle thermal properties in the thermal con-
to the thermal properties of the Al-based composites. ducting behavior of Al-based nanocomposites. Figure 10

250
Thermal conductivity (W/mK)

220

190

160
With ITR
130
Without ITR
100
150 170 190 210 230 250
Metal matrix thermal conductivity (W/mK)

Fig. 9  Influence of metal matrix thermal conductivity on the thermal


Fig. 7  Influence of degree of nanoparticle agglomeration on the ther- conductivity of Al-based nanocomposites reinforced with nanoscale
mal conductivity of Al-based nanocomposites SiC particles

13
Arabian Journal for Science and Engineering

distribution of nanoparticles maximized the thermal conduc-


Normalized PMMNC thermal conductivity tivity. Thus, it was suggested that producing a homogenous
1.2
PMMNC is another way to improve the thermal properties.
The results depicted that the Al-based nanocomposite ther-
1.14
mal conductivity enhances by increasing both amount and
size of nanoscale ceramic particles. Moreover, the effective
1.08
thermal conducting coefficients of the PMMNCs can be
greatly improved with rising the thermal conductivities of
1.02 5%, Without ITR the metallic matrix and ceramic nanoparticles. The devel-
10%, Without ITR oped methodology may be a suitable model to design the
0.96 5%, With ITR metallic composites with high thermal conductivity utilized
10%, With ITR in micro-electro-mechanical systems.
0.9
1 2.5 4 5.5 7 8.5 10
Normalized nanoparticle thermal conductivity

Fig. 10  Variation of normalized thermal conductivity of SiC nano-


References
particle-reinforced Al-based nanocomposites with normalized ther-
mal conductivity of ceramic nanoparticle 1. Niamat, M.; Sarfraz, S.; Shehab, E.; Ismail, S.O.; Khalid, Q.S.:
Experimental characterization of electrical discharge machining
of aluminum 6061 T6 alloy using different dielectrics. Arabian J.
Sci. Eng. 44(9), 8043–8052 (2019)
displays the variation of the normalized thermal conductiv- 2. Hassanzadeh-Aghdam, M.K.: Micromechanics-based thermal
ity of PMMNC with the normalized thermal conductivity expansion characterization of SiC nanoparticle-reinforced metal
matrix nanocomposites. Proc. Inst. Mech. Eng. Part C J. Mech.
of ceramic nanoparticle. The micromechanical simulation
Eng. Sci. 233(1), 190–201 (2019)
has been performed for two weight fractions (5% and 10%). 3. Varol, T.; Canakci, A.; Yalcin, E.D.: Fabrication of nanoSiC-rein-
Too, the role of ITR in the PMMNC thermal conductivity forced Al2024 matrix composites by a novel production method.
has been tested. Note that both thermal conductivities are Arabian J. Sci. Eng. 42(5), 1751–1764 (2017)
4. Chen, X.; Fu, D.; Teng, J.; Zhang, H.: Hot deformation behavior
normalized with respect to the matrix thermal conductiv-
and mechanism of hybrid aluminum-matrix composites reinforced
ity. Generally, a nonlinear increase in the PMMNC thermal with micro-SiC and nano-TiB2. J. Alloy. Compd. 753, 566–575
conductivity is found by increasing the ceramic nanoparti- (2018)
cle thermal conductivity. The effect of nanoscale particle 5. Lakshmikanthan, A.; Bontha, S.; Krishna, M.; Koppad, P.G.;
Ramprabhu, T.: Microstructure, mechanical and wear properties
thermal properties is more pronounced at a higher weight
of the A357 composites reinforced with dual sized SiC particles.
fraction. J. Alloy. Compd. 786, 570–580 (2019)
6. Fei, W.D.; Hu, M.; Yao, C.K.: Thermal expansion and thermal
mismatch stress relaxation behaviors of SiC whisker reinforced
aluminum composite. Mater. Chem. Phys. 77(3), 882–888 (2003)
4 Conclusions 7. Sahu, S.; Mondal, D.P.; Cho, J.U.; Goel, M.D.; Ansari, M.Z.:
Low-velocity impact characteristics of closed cell AA2014-SiCp
In this work, a micromechanics-based hierarchical tech- composite foam. Compos. B Eng. 160, 394–401 (2019)
nique involving the M-T method and the modified effective 8. Xu, Y.; Tanaka, Y.; Goto, M.; Zhou, Y.; Yagi, K.: Thermal con-
ductivity of SiC fine particles reinforced al alloy matrix composite
medium approach was developed to estimate the effective with dispersed particle size. J. Appl. Phys. 95(2), 722–726 (2004)
thermal conductivities of Al-based nanocomposites filled 9. Safi, M.; Hassanzadeh-Aghdam, M.K.; Mahmoodi, M.J.: Effects
with the SiC nanoparticles. The effects of the ITR between of nano-sized ceramic particles on the coefficients of thermal
the nanoparticle and metallic matrix as well as the disper- expansion of short SiC fiber-aluminum hybrid composites. J.
Alloy. Compd. 803, 554–564 (2019)
sion type of nanoscale particles were examined. When the 10. Suryanarayana, C.; Al-Aqeeli, N.: Mechanically alloyed nanocom-
agglomeration and ITR were included in the microme- posites. Prog. Mater Sci. 58(4), 383–502 (2013)
chanical analysis, a good agreement was found between 11. Keneshloo, M.; Paidar, M.; Taheri, M.: Role of SiC ceramic parti-
the predictions and the experimental results. Generally, the cles on the physical and mechanical properties of Al–4% Cu metal
matrix composite fabricated via mechanical alloying. J. Compos.
nanoscale SiC particles lead to an insignificant enhancement Mater. 51(9), 1285–1298 (2017)
in the Al-based nanocomposite thermal conductivities which 12. Casati, R.; Fiocchi, J.; Fabrizi, A.; Lecis, N.O.R.A.; Bonollo,
was attributed to the presence of ITR. It was offered that F.; Vedani, M.: Effect of ball milling on the ageing response of
one well-organized way to improve the PMMNC thermal Al2618 composites reinforced with SiC and oxide nanoparticles.
J. Alloy. Compd. 693, 909–920 (2017)
conductivity is to reduce the interfacial resistance. Also, 13. del Rio, E.; Nash, J.M.; Williams, J.C.; Breslin, M.C.; Daehn,
the ceramic nanoparticle agglomeration decreased the Al- G.S.: Co-continuous composites for high-temperature applica-
based nanocomposite thermal properties, while a uniform tions. Mater. Sci. Eng., A 463(1–2), 115–121 (2007)

13
Arabian Journal for Science and Engineering

14. Han, J.; Hong, C.; Zhang, X.; Wang, B.: Thermal shock resist- 28. Chang, J.; Zhang, Q.; Lin, Y.; Wu, G.: Layer by layer graphite film
ance of T­ iB2–Cu interpenetrating phase composites. Compos. Sci. reinforced aluminum composites with an enhanced performance
Technol. 65(11–12), 1711–1718 (2005) of thermal conduction in the thermal management applications.
15. Zhang, X.; Hong, C.; Han, J.; Zhang, H.: Microstructure and J. Alloy. Compd. 742, 601–609 (2018)
mechanical properties of ­TiB2/(Cu, Ni) interpenetrating phase 29. El-Kady, O.; Fathy, A.: Effect of SiC particle size on the physical
composites. Scripta Mater. 55(6), 565–568 (2006) and mechanical properties of extruded Al matrix nanocomposites.
16. Kollo, L.; Bradbury, C.R.; Veinthal, R.; Jäggi, C.; Carreno- Mater. Des. 54, 348–353 (2014)
Morelli, E.; Leparoux, M.: Nano-silicon carbide reinforced alu- 30. Nan, C.W.; Birringer, R.; Clarke, D.R.; Gleiter, H.: Effective
minium produced by high-energy milling and hot consolidation. thermal conductivity of particulate composites with interfacial
Mater. Sci. Eng., A 528(21), 6606–6615 (2011) thermal resistance. J. Appl. Phys. 81(10), 6692–6699 (1997)
17. Murthy, N. V., Reddy, A. P., Selvaraj, N., & Rao, C. S. P. (2016, 31. Huang, C.L.; Qian, X.; Yang, R.G.: Influence of nanoparticle size
September).: Preparation of SiC based Aluminium metal matrix distribution on the thermal conductivity of particulate nanocom-
nano composites by high intensity ultrasonic cavitation process posites. EPL (Europhys. Lett.) 117(2), 24001 (2017)
and evaluation of mechanical and tribological properties. In IOP 32. Chu, K.; Wu, Q.; Jia, C.; Liang, X.; Nie, J.; Tian, W.; Guo, H.:
Conference Series: Materials Science and Engineering (Vol. 149, Fabrication and effective thermal conductivity of multi-walled
No. 1, p. 012106). IOP Publishing. carbon nanotubes reinforced Cu matrix composites for heat sink
18. Mazahery, A.; Shabani, M.O.: Characterization of cast A356 alloy applications. Compos. Sci. Technol. 70(2), 298–304 (2010)
reinforced with nano SiC composites. Transac. Nonferrous Met. 33. Hemant, P.A.L.; Sharma, V.: Thermal conductivity of carbon
Soc. China 22(2), 275–280 (2012) nanotube-silver composite. Transac. Nonferrous Met. Soc. China
19. El-Daly, A.A.; Abdelhameed, M.; Hashish, M.; Eid, A.M.: Syn- 25(1), 154–161 (2015)
thesis of Al/SiC nanocomposite and evaluation of its mechanical 34. Liu, H.; Zhou, S.; Li, X.: Inferring the size distribution of 3d
properties using pulse echo overlap method. J. Alloy. Compd. 542, particle clusters in metal matrix nanocomposites. J. Manuf. Sci.
51–58 (2012) Eng. 135(1), 011013 (2013)
20. Abdullahi, K.; Al-Aqeeli, N.: Mechanical alloying and spark 35. Dastgerdi, J.N.; Marquis, G.; Salimi, M.: The effect of nanotubes
plasma sintering of nano-SiC reinforced Al–12Si–0.3 Mg alloy. waviness on mechanical properties of CNT/SMP composites.
Arabian J. Sci. Eng. 39(4), 3161–3168 (2014) Compos. Sci. Technol. 86, 164–169 (2013)
21. Boostani, A.F.; Tahamtan, S.; Jiang, Z.Y.; Wei, D.; Yazdani, S.; 36. Yang, Q.S.; He, X.Q.; Liu, X.; Leng, F.F.; Mai, Y.W.: The effec-
Khosroshahi, R.A.; Gong, D.: Enhanced tensile properties of alu- tive properties and local aggregation effect of CNT/SMP compos-
minium matrix composites reinforced with graphene encapsulated ites. Compos. B Eng. 43(1), 33–38 (2012)
SiC nanoparticles. Compos. A Appl. Sci. Manuf. 68, 155–163 37. Nemat-Nasser, S.; Hori, M.: Micromechanics: overall properties
(2015) of heterogeneous materials. Elsevier, London (2013)
22. Yang, Y.; Lan, J.; Li, X.: Study on bulk aluminum matrix nano- 38. Mori, T.; Tanaka, K.: Average stress in matrix and average elastic
composite fabricated by ultrasonic dispersion of nano-sized SiC energy of materials with misfitting inclusions. Acta Metall. 21(5),
particles in molten aluminum alloy. Mater. Sci. Eng., A 380(1–2), 571–574 (1973)
378–383 (2004) 39. Benveniste, Y.: A new approach to the application of Mori-Tana-
23. Casati, R.; Wei, X.; Xia, K.; Dellasega, D.; Tuissi, A.; Villa, E.; ka’s theory in composite materials. Mech. Mater. 6(2), 147–157
Vedani, M.: Mechanical and functional properties of ultrafine (1987)
grained Al wires reinforced by nano-Al2O3 particles. Mater. Des. 40. Chen, C.H.; Wang, Y.C.: Effective thermal conductivity of misori-
64, 102–109 (2014) ented short-fiber reinforced thermoplastics. Mech. Mater. 23(3),
24. Zhou, D.; Qiu, F.; Jiang, Q.: The nano-sized TiC particle rein- 217–228 (1996)
forced Al–Cu matrix composite with superior tensile ductility. 41. Hatta, H.; Taya, M.: Effective thermal conductivity of a misori-
Mater. Sci. Eng., A 622, 189–193 (2015) ented short fiber composite. J. Appl. Phys. 58(7), 2478–2486
25. Miranda, A.; Barekar, N.; McKay, B.J.: MWCNTs and their use (1985)
in Al-MMCs for ultra-high thermal conductivity applications: a 42. Nan, C.W.; Liu, G.; Lin, Y.; Li, M.: Interface effect on thermal
review. J. Alloy. Compd. 774, 820–840 (2019) conductivity of carbon nanotube composites. Appl. Phys. Lett.
26. Hasselman, D.P.H.; Donaldson, K.Y.; Geiger, A.L.: Effect of rein- 85(16), 3549–3551 (2004)
forcement particle size on the thermal conductivity of a partic- 43. Mortazavi, B.; Bardon, J.; Ahzi, S.: Interphase effect on the elas-
ulate-silicon carbide-reinforced aluminum matrix composite. J. tic and thermal conductivity response of polymer nanocompos-
Am. Ceram. Soc. 75(11), 3137–3140 (1992) ite materials: 3D finite element study. Comput. Mater. Sci. 69,
27. Molina, J.M.; Rhême, M.; Carron, J.; Weber, L.: Thermal conduc- 100–106 (2013)
tivity of aluminum matrix composites reinforced with mixtures of
diamond and SiC particles. Scripta Mater. 58(5), 393–396 (2008)

13

You might also like