You are on page 1of 15

International Journal of Thermal Sciences 171 (2022) 107209

Contents lists available at ScienceDirect

International Journal of Thermal Sciences


journal homepage: www.elsevier.com/locate/ijts

Influence of graphene nano-platelets on thermal transport performance of


carbon fiber-polymer hybrid composites: Overall assessment of
microstructural aspects
Mohammad Kazem Hassanzadeh-Aghdam a, *, Reza Ansari b, Hamed Mohaddes Deylami a
a
Faculty of Technology and Engineering, East of Guilan, University of Guilan, Rudsar, Iran
b
Faculty of Mechanical Engineering, University of Guilan, Rasht, Iran

A R T I C L E I N F O A B S T R A C T

Keywords: This study comprehensively investigates the thermal transport behavior of unidirectional hybrid composites
Hybrid composite (UHCs) by means of a micromechanical methodology. The constructional feature of the UHC is that the
Thermal transport performance continuous carbon fibers are embedded in the graphene nano-platelets (GNPs)-modified epoxy resin. The in­
Nano-graphene
fluences of volume fraction, length, thickness, and alignment of GNPs, the interfacial thermal resistance (ITR)
Agglomeration
Micromechanics
between the nano-graphene and polymer matrix, as well as the volume fraction, arrangement type, and off-axis
angle of carbon fibers on the thermal conductivities of UHCs are extensively analyzed. Further, the micro­
mechanical model is extended to account the effect of GNP agglomeration on the UHC thermal conductivities.
The results show that uniformly dispersed GNPs play a dominant role in improving the UHC thermal conductivity
along the transverse direction, while the axial thermal conductivity is insignificantly influenced by the nano-
graphene particles. Besides, using the GNPs with a higher aspect ratio (length/thickness) is an efficient
manner to obtain much better thermal transport performance for the UHCs. It is observed that the formation of
GNP agglomeration within the epoxy resin severely decreases the transverse thermal conductivity. The presence
of GNP/epoxy ITR is a lowering factor of the thermal conductivity. As compared to the hexagonal and random
arrays, the square array of carbon fibers within the GNP-modified epoxy produces the largest transverse thermal
conductivity. On the basis of comparative studies, the model predictions agree very well with the experimental
data available in the literature.

1. Introduction fiber fracture, matrix yielding, fiber/matrix interfacial debonding, as


well as the delamination in the laminated composites. Since the local­
Carbon fiber reinforced polymer composites (CFRPCs) have been ized damages can result in catastrophic failure of the CFRPC structures
extensively used as primary structures in aerospace, automotive, and [7], the accumulated heat must be dissipated. Therefore, development
renewable energy industries because of their high specific strength and of CFRPCs with high thermal transport performance to hinder incipient
stiffness, remarkable fatigue characteristics and good corrosion resis­ thermal-induced structural damage is of great interest to numerous in­
tance [1–3]. The CFRPCs have excellent in-plane properties dominated dustrial sectors.
by the high-performance carbon fibers. But, the out-of-plane properties Tremendous efforts have been devoted to improve the out-of-plane
of CFRPCs governed by the matrix are very low which can limit their thermal conductivities of fiber/polymer composites. It was stated that
structural applications [4]. For instance, the low out-of-plane thermal the out-of-plane heat transfer mechanisms in CFRPCs are intensely
conductivity of CFRPCs is due to the poor thermal conductivity of the affected by the thermal conductivity of the matrix. It seems that adding
polymer matrixes [5,6]. Generally, the directional conduction can be an thermally conductive nano-sized carbon particles into the polymer
advantage of fibrous composite materials that can be conducting in one matrix to be a practical method to achieve suitable out-of-plane heat
direction and insulating in the other. Consequently, heat is not effi­ dissipation performance [7–9]. The main reason is that carbonaceous
ciently conducted away which may lead to several damage types, such as nanomaterials efficiently improve the matrix thermal transport

* Corresponding author.
E-mail address: Mk.hassanzadeh@gmail.com (M.K. Hassanzadeh-Aghdam).

https://doi.org/10.1016/j.ijthermalsci.2021.107209
Received 29 September 2020; Received in revised form 30 June 2021; Accepted 2 August 2021
Available online 10 August 2021
1290-0729/© 2021 Elsevier Masson SAS. All rights reserved.
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 1. Schematic sketch of the GNP/fiber-reinforced UHCs.

coefficient [10,11]. Among carbon-based particles, graphene consisting additive was 53.7 % greater than that of the carbon fiber/PEEK com­
of a sheet of carbon atoms arranged in a honeycomb structure [11,12] posite [19]. According to the test results of Noh and Kim [20], the
offers the potential to enhance the thermal properties of polymer com­ isotropic thermal conductivity of the 5 wt% carbon fiber/15 wt%
posites [11,13]. It is due to the ultrahigh thermal conductivity of gra­ GNP-reinforced hybrid composite attained improvement of 82 % as
phene (~5300 W/mK) [14]. Graphene nano-platelets (GNPs) are unique compared to the 20 wt% GNP-reinforced composite, and improvement
nanoparticles composed of several layers of graphene. Due to the low of 65 % as compared to the 20 wt% carbon fiber-reinforced composite.
cost, the mass production of GNPs is suitable [15]. The addition of GNPs Heat transfer mechanism of fiber/GNP/polymer hybrid composites is
into the fiber/polymer composites promises an efficient and economical largely influenced by the amount, size, and dispersion degree of GNPs,
route to create a new class of hybrid composites for thermal manage­ volume fraction, and orientation of carbon fibers and most importantly
ment in high-tech industries such as aircraft structural applications. As the nanofiller/polymer interfacial thermal resistance (ITR) [7,9,11,13].
compared to the conventional fiber/polymer composites, the fiber/­ Up to now, no systematic study concerning the overall thermal con­
GNP/polymer hybrid composites combine the properties of all the ductivities of unidirectional hybrid composites (UHCs) consisting of the
constituents and display improved performance. carbon fiber, GNP and epoxy resin has been performed. It is imperative
A number of research works have been devoted to analyze the to develop innovative strategies for predicting the thermal conductiv­
thermal conductivity of carbon fiber/polymer hybrid composites con­ ities of UHCs and investigate the effects of critical microstructural as­
taining graphene particles [7,16–18]. For instance, Kandare et al. [7] pects on heat transfer performance.
found that the inclusion of 1 vol% GNPs into the carbon fiber/epoxy In addition to experimental methods, a lot of simulation tools have
composites results in 9 % enhancement in the out-of-plane thermal been designed to predict the effective thermal conductivities of com­
conductivity of the hybrid composites at room temperature. Wang and posite materials [13,21–23]. The main advantages of the theoretical
Cai [9] experimentally investigated the effect of GNP weight fraction on methods are their simplicity and low cost. In this work, a hierarchical
the thermal conductivity of the carbon fiber/epoxy hybrid composites. micromechanical method is proposed to analyze the thermal conduc­
The out-of-plane thermal conductivity of the hybrid composite con­ tivity of carbon fiber/GNP/epoxy UHCs. The organizations of the pre­
taining 0.5 wt% GNPs indicated a 55.6 % enhancement as compared to sent research paper are as follows. The micromechanical models and
the base carbon/epoxy composite [9]. Zhang et al. [18] produced the procedures of predicting the UHC thermal conductivities are established
epoxy nanocomposites modified with 0.5, 1, 2, and 4 wt% graphene in Section 2. Section 3 presents the numerical results to investigate the
nanoflakes. These nanocomposite materials were used as the matrix effects of volume fraction, length, thickness, agglomeration degree, and
phase for carbon fibers to manufacture the hybrid composites. The alignment of GNPs, the ITR between the nano-graphene and polymer
hybrid composite thermal properties were improved with increasing the matrix, volume fraction, arrangement type, and off-axis angle of carbon
graphene content. The thermal conductivity of the carbon fiber/epoxy fibers on the UHC thermal transport properties. Besides, the simulation
hybrid composites with 4 % nano-graphene was improved 45 % as results are validated by comparing them with the experimental data
compared to the base carbon fiber/epoxy composites [18]. By ball mainly to confirm the ability of micromechanical methodology in pre­
milling and hot-press processing, Liu et al. [19] produced carbon dicting the thermal conductivities.
fiber-reinforced poly(ether ether ketone) (PEEK) hybrid composites with
different weight fractions of graphene. The thermal conductivity of the
carbon fiber/PEEK hybrid composite containing 1 wt% graphene

Fig. 2. Schematic representation of (a) GNP as an equivalent oblate spheroid and (b) GNP coated with a thin interfacial thermal barrier layer.

2
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

2. Theoretical approach

The thermal conductivity of GNP-modified polymer matrix nano­


composites (PMNCs) is obtained within the framework of the micro­
mechanics first. The important microstructural aspects of PMNCs,
particularly the nanoparticle/polymer ITR and the GNP agglomeration
are taken into account. Then, a unit cell-based micromechanical model
is developed to estimate the thermal conductivities of unidirectional
fiber-reinforced composites. Finally, the detailed procedures of the
prediction of UHC thermal transport coefficients are presented. Fig. 1
shows a schematic sketch of the GNP/fiber-reinforced UHCs.

2.1. GNP-modified PMNCs

The GNPs are assumed to be uniformly distributed and randomly


oriented in the polymer matrix. The GNP is modeled as an equivalent
oblate spheroid with a specified aspect ratio p (length/thickness)
[24–26]. The review of literature confirms that the ITR (RK ) between the
nanoparticles and polymer matrix has a high influence on the thermal
transport behavior of composite materials [27–30]. The nano­
particle/polymer ITR denotes all combined effects which contribute to
an interfacial thermal barrier [24,31]. To take the ITR influence into Fig. 3. Schematic sketch of the PMNC containing GNP agglomeration.
account, the nano-graphene particle is seen as a nanoscale plate coated
with a very thin interfacial thermal barrier layer [24,26,32], as shown in GNP-reinforced nanocomposites are extracted. In Eq. (3), H reflects the
Fig. 2. In this figure l and t are the length and thickness of GNP. The geometrical factors and is expressed as [24].
equivalent thermal conductivities of GNP along the in-plane (Kx ) and ⎡ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ⎤
through-thickness (Kz ) directions are given by [24]. ( )2
ln⎣tl + l
t
− 1 ⎦ tl
Kg Kg
Kx = 2RK Kg , Kz = 2RK Kg (1) H= √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
1
(4)
l
+ 1 t
+1 [( )2 ]3̅ − ( )2 .
l
l
− 1 t
− 1
t
where Kg is the GNP thermal conductivity. The thickness of GNP
robustly affects its thermal conductivity [33,34]. The thickness depen­ It has been previously mentioned that the agglomeration of nano-
dence of Kg can be expressed as follows [34]. graphene particles is one of the most significant problems decreasing
( )− 1/2 the graphene capability in improving the effective properties of PMNCs
t
Kg = 4058 × 10
. (2) [36–39]. Herein, a micromechanics-based technique is offered in a
3.4 × 10−
nested modeling procedure to capture the GNP agglomeration. Such
Thus, the magnitude of Kg can be determined from the thickness of method has been already used to investigate the role of nanotube
GNP. Micromechanical methods are able to predict the thermal con­ agglomeration in the elastic properties of polymer composite materials
ductivity of the GNP-filled composites. Several assumptions and limi­ [40–42]. In this approach, a number of GNPs are uniformly dispersed
tations exist in these methods. For example, the micromechanical inside the polymer matrix making the equivalent matrix, and other GNPs
models have been fundamentally developed for the continuum rein­ appear in the agglomeration form, as presented in Fig. 3. It is assumed
forcing agent, while graphene has a lattice structure. Furthermore, a that the regions with the agglomerated GNPs have spherical shapes, and
perfect structure without any defects is assumed for GNPs. In the are incorporated as “inclusions”. The whole volume of GNPs in the
micromechanical modeling, a similar size is supposed for all GNPs PMNC, signified by VGNP , can be divided into two parts as
dispersed within the composites. The interfacial area between nano- SI
VGNP = VGNP EM
+ VGNP (5)
graphene and polymer is governed by the molecular interactions,
while an equivalent solid continuum interphase between the nano­ SI EM
where VGNP is the GNP volume in the spherical inclusions, and VGNP is
particle and polymer is frequently assumed in the micromechanical
the GNP volume in the effective matrix. Two parameters to describe the
rules. Besides, formation of voids is not considered in the micro­
GNP agglomeration are presented as
mechanical modeling of the thermal conductivity of the GNP-reinforced
composites. On the basis of micromechanics approach [24,34,35], the VSI V SI
ξ= , ζ = GNP 0 ≤ ξ, ζ ≤ 1 (6)
effective thermal conductivity of the PMNCs with randomly oriented V VGNP
GNPs (KPMNC ) can be predicted by the following relation
⎡ ⎤ ⎡ ⎤ where V and VSI are the PMNC volume and the spherical inclusion
volume in the PMNCs, respectively. ξ specifies the volume fraction of
2⎢ fGNP KM ⎥ 1⎢ fGNP KM ⎥
KPMNC = KM + ⎢ ( )− 1 ⎥ ⎢
⎦ +3⎣ ( )− 1 ⎥ (3) inclusions with respect to the PMNC volume, and ζ denotes the volume
3⎣ ⎦
H + KKMx − 1 1− H
+ KKMz − 1 ratio of GNPs inside the spherical inclusions over the whole GNP inside
the PMNC. A uniform dispersion of GNPs throughout the PMNCs
2

without agglomeration occurs when ξ = 1. By the reduction of ξ, the


where KM denotes the thermal conductivity of the polymer matrix, and degree of GNP agglomeration becomes more severe. Once ζ = 1, all the
fGNP signifies the GNP volume fraction. The corresponding formulas for GNPs are located in the sphere zones. The case ξ = ζ means that all the
the axial and transverse thermal conductivities of the aligned GNP- GNPs are dispersed evenly without agglomeration. By the consideration
reinforced composites can be derived by means of Eq. (3) as intro­ of fGNP ( = VGNP /V) along with Eqs. (5) and (6), following relations can
duced in Ref. [35]. Thus, the results about the orientation of GNPs can be be derived as
investigated. In other words, the thermal conductivities of the aligned

3
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 4. Schematic representation of an RVE with a square array packing of fibers.

Fig. 5. Extension of the RVE of the SUCM

SI
VGNP ζ V EM (1 − ζ) Several assumptions are imposed in developing Eq. (8). For example,
SI
fGNP = EM
= fGNP , fGNP = GNP = fGNP (7)
VSI ξ V − VSI (1 − ξ) the spherical inclusion in the representative volume element (RVE) is
replaced by a cubic inclusion of the same volume so as to keep the same
SI
where fGNP EM
and fGNP signify the GNP volume fractions in the spherical volume fraction. Furthermore, spherical inclusions with similar shapes
inclusions and in the equivalent matrix, respectively. Therefore, the and sizes are dispersed uniformly in the matrix. Both phases (inclusion
total GNP-modified PMNC is a composite material consisted of the and matrix) behave as isotropic materials. Additionally, a perfectly
spherical inclusions embedded in the equivalent matrix. The micro­ bonded interface between the inclusion and the matrix is assumed, and
mechanical modeling starts from the prediction of the thermal con­ no interfacial resistance between these two phases is considered.
ductivities of spherical inclusion (KSI ) and equivalent matrix (KEM ) using
Agg
Eq. (3). Then, we calculate KPMNC , the effective thermal conductivity of 2.2. Unidirectional composites
the whole PMNC with the GNP agglomeration by capturing the spherical
inclusion as the reinforcement and the equivalent matrix as the new The micromechanical method based on the simplified unit cell model
matrix, as follows [43]. (SUCM) is presented to evaluate the thermal transport behavior of
unidirectional fiber-reinforced composites. The SUCM has been exten­
ξ0.67 KEM ( )
Agg
KPMNC = ( ) + 1 − ξ0.67 KEM . (8) sively used to predict the mechanical properties of unidirectional com­
1 − ξ0.33 1 − posites [44,45]. The SUCM with an RVE consisted of four sub-cells has
KEM
KSI
been developed to estimate the thermal conductivities of unidirectional

4
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

composites [46]. The cross-section of composite was idealized with a


square regular array of fibers [46], as shown in Fig. 4. However, the real
arrangement of the fibers within the cross-section of the unidirectional
composite is random. Therefore, the micromechanical formulations of
the SUCM should be modified to simulate the random array of fibers
embedded in the matrix [8]. The geometry of the RVE of the SUCM is
extended to r × c sub-cells, as can be seen in Fig. 5. The size of each
sub-cell and whole RVE can be seen in this figure. The perfect bonding
conditions are applied between the RVE sub-cells.
Introducing the global heat flow Ql (l = x, y, z) over the RVE leads to
the following equilibrium equations for the heat flux within the sub-cells
as [46].

r ∑
c r ∑
∑ c
bj q1j
x = Qx L r , ai qi1
y = Qy Lc , bj ai qijz = Qz Lr Lc . (9)
j=1 i=1 j=1 i=1

ij
where ql is the local heat flux in the sub-cell ij in direction l. According to
the equivalence of the local heat flux components at the sub-cell in­
terfaces, following relations are extracted as [8,46,47].

q1j ij i1 ij
x = qx (i > 1), qy = qy (j > 1). (10)

In the unit cell approaches, the temperature gradient inside the sub-
cells of the RVE is supposed to be linear [46,47]. The temperature
compatibility on the basis of RVE description leads to


⎪ ∑ c ∑c

⎪ ai T,xi1 = ai T,xij = Lc T ,x (j > 1),



⎪ i=1 i=1

∑ ∑ (11)
r r

⎪ bj T,y1j = bj T,yij = Lr T ,y (i > 1) Fig. 6. Multi-step modeling of the thermal conductivities of UHCs.

⎪ j=1

⎪ j=1


⎩ T ij = T (i > 1, j > 1) 2.3. Thermal conductivities of UHCs
,z ,z

ij
in which T,l is the local temperature gradient inside the sub-cell ij, and T,l Fig. 6 shows the various steps involved in the micromechanical
prediction of the UHC thermal conductivities. First, the thermal con­
is the global temperature gradient in direction l. On the basis of Fourier’s
ductivities of the spherical inclusion and equivalent matrix are needed.
law, the relation between the heat flow vector q and temperature
By considering ξ, ζ, fGNP , and the properties of GNP and polymer, Eq. (3)
gradient for the sub-cell ij is expressed as
is employed to calculate the values of KSI and KEM . These results are used
Tij, = Kij
− 1
: qij (12) as an input to the micromechanical model in Eq. (8) to estimate the
thermal conductivity of GNP-modified PMNCs. In this step, the spherical
where Kij stands for the thermal conductivity tensor of the sub-cell ij. A inclusion plays the role of reinforcing phase and the equivalent matrix is
system of cr + c + r equations with the identical number of unknowns considered as the new matrix phase. The combination of the fibers and
will be obtained by substituting Eq. (12) in Eq. (11) along with Eqs. (9) the PMNC is viewed as a UHC containing GNPs that its thermal con­
and (10), as follows ductivity is estimated by the SUCM. In the RVE of UHCs, the rein­
forcement is the unidirectional fiber and the matrix phase is the GNP-
Am×m qm×1 = Fm×1 where m = cr + c + r (13) modified PMNCs.

in which q and F denote the heat flux vector inside each sub-cell and the
3. Results and discussion
external heat flux vector, respectively. Moreover, A specifies the coef­
ficient tensor made of the geometrical and material characteristics of the
To verify the validity of the micromechanical model in estimating the
sub-cells. Solving Eq. (13) gives the components of local heat flux for all
thermal conductivity of the GNP-modified PMNCs, the calculations are
sub-cells. Then, the components of local temperature gradient for all
compared with the experimental data [48]. Fig. 7 shows the comparison
sub-cells are calculated by Eq. (12). The components of global temper­
between two sets of results. The thermal conductivity of the GNP-epoxy
ature gradient are determined by Eq. (11). Finally, the thermal con­
nanocomposite is plotted as a function of GNP volume fraction. Two
ductivities in directions z, y and x are estimated, respectively, as
types of GNPs, including (i) multi-layered GNP with l = 1.7 μm and t =
Qz 60 nm, and (ii) few-layered GNP with l = 0.35 μm and t = 1.7 nm are
(14)
considered [34,48]. The value of RK is 3.5 × 10− 9 m2K/W [34,49]. The
Kz = when Qz ∕
= 0 and Qx = Qy = 0,
T ,z
micromechanical modeling is performed for two conditions of nano­
Qy particle dispersion, including (i) uniformly dispersed GNPs, and (ii)
Ky = when Qy ∕
= 0 and Qx = Qz = 0, (15)
T ,y agglomerated GNPs with ξ = 0.5 and ζ = 0.85. It is found from Fig. 7 a
and b that a substantial discrepancy exists between the experimental
Qx outcomes [48] and the predicted values by the model with uniformly
Kx = when Qx ∕
= 0 and Qy = Qz = 0. (16)
T ,x dispersed GNPs and ignoring the GNP/epoxy ITR. This is more obvious
at higher GNP volume fraction. Considering only the ITR in the micro­
It is noticed that for a transversely isotropic composite, we have Kx =
mechanical modeling leads to the predictions to be close to the experi­
Ky =
∕ Kz .
mental measurements at the low graphene content. Nevertheless, the

5
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 7. Comparison of the model predictions and experimental measurements [48] of the thermal conductivity of the GNP-modified epoxy nanocomposites, (a) l =
1.7 μm and t = 60 nm, and (b) l = 0.35 μm and t = 1.7 nm.

Fig. 8. Effect of agglomeration parameter on the thermal conductivity of the GNP-modified epoxy nanocomposites, (a) l = 1.7 μm and t = 60 nm, and (b) l = 0.35
μm and t = 1.7 nm.

6
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

and experimental measurements, the values of ξ and ζ should be about


0.5 and 0.85, respectively.
The effect of ITR values on the thermal conductivity of GNP/epoxy
nanocomposites is investigated as shown in Fig. 9. When the value of ITR
is 3.5 × 10− 9 m2K/W, a good agreement exists between the predictions
and experiments [48]. The thermal conductivity of the polymer nano­
composites decreases as the value of ITR increases. A great amount of
phonon scattering can be occurred at the filler/matrix interface because
of the unbalanced density of phonon states. This causes high thermal
resistance during the interfacial thermal transport process [7].
The effects of geometrical characteristics and dispersion type of
GNPs on the thermal conductivity of the epoxy nanocomposites are
investigated. Both uniform (or perfect) dispersion and alignment of
GNPs are considered. This case may be an examination about the role of
GNP orientation in the nanocomposite thermal conductivity. To analyze
in more detail, the micromechanical modeling is accomplished under
two conditions; (i) with ITR, and (ii) without ITR. The variation of the
nanocomposite thermal conductivity with the GNP length is depicted in
Fig. 10a. It is found that the thermal conductivity of the epoxy nano­
composites in the presence of ITR is nonlinearly enhanced by increasing
the GNP length. More thermal conductive paths can be constructed by
the increase of GNP length which facilitates the heat transfer. The results
Fig. 9. Effect of ITR value on the thermal conductivity of the GNP-modified
show that the thermal conductivity of the aligned GNP-reinforced
epoxy nanocomposites.
nanocomposites is higher than that of the randomly dispersed GNP-
reinforced nanocomposites. Thus, a better thermal transport can be
thermal conductivities achieved by the model considering the ITR and conducted by the alignment of GNPs along the thermal loading. When
uniform dispersion of GNPs are still far from the experiments at the the ITR is not taken into account in the modeling, the thermal conduc­
higher GNP content. Excellent agreement between two sets of results tivity of the epoxy nanocomposite suddenly increases by the increase of
occurs when both the agglomeration of GNPs and graphene/polymer GNP length up to 0.75 μm and then, its value saturates. In other words,
ITR are taken into account in the micromechanical analysis. As expected when l > 0.75 μm, further increase of GNP length does not affect the
from the properties of graphene, the thermal conductivity of the epoxy nanocomposite thermal conductivity. Furthermore, in the absence of
nanocomposites increases with the increase of nanoparticle volume ITR, the alignment of GNPs leads to higher thermal conductivities for the
fraction. Based on the micromechanical results, the ITR between the polymer nanocomposites as compared to the random dispersion. The
GNP and epoxy matrix significantly decreases the thermal properties of thermal conductivity of GNP-epoxy nanocomposites as a function of
the PMNCs. Because of the discrepancy in the phonon density of state, GNP thickness is shown in Fig. 10b. With increasing the thickness of
phonons are scattered at the GNP/epoxy interface and thus, the thermal GNP up to about 40 nm, the thermal conductivity decreases. In the
transport is seriously hindered. The PMNCs filled with uniformly absence of ITR, the thermal conductivity drop is very notable. When t >
dispersed GNPs have a higher thermal conductivity than those filled 40 nm, variation of GNP thickness insignificantly affects the nano­
with non-uniformly dispersed GNPs; i.e. agglomerated state. Thus, GNPs composite thermal conductivity. The thermal conductivity is further
with a homogenous dispersion are really beneficial to the heat transfer enhanced by aligning the GNPs in the polymer matrix. It can be
ability of the PMNCs. concluded from Fig. 10 that the variations of GNP orientation and shape
The micromechanical predictions of the thermal conductivity of the change the thermal conductivity of the polymer nanocomposites
GNP/epoxy nanocomposite with different values of agglomeration without (or with) the ITR.
parameter are provided in Fig. 8. Both nanocomposite systems with Fig. 11 shows the role of GNP thermal conductivity on the thermal
different dimensional characteristics of GNPs are analyzed. The values conductivity of the GNP-modified epoxy nanocomposites. The theoret­
of ξ and RK are 0.85 and 3.5 × 10− 9 m2K/W, respectively. Also, the ical results are provided under two conditions for GNP thermal con­
experimental measurements [48] are given in this figure. The results ductivity; (i) Kg = 5300 W/mK, and (ii) Kg based on Eq. (2) which is a
indicate that to have the best agreement between the model predictions

Fig. 10. Variation of the thermal conductivity of the GNP-modified epoxy nanocomposites with (a) GNP length, t = 1.7 nm, (b) GNP thickness, l = 0.35 μm.

7
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

1 1 − fF fF
= + . (20)
Klb KM KF
As can be seen in Fig. 13a, three sets of results predicted by the
SUCM, CCA, upper bond are in excellent agreement. It is found from
Fig. 13b that a good agreement exists between the results of the SUCM
and CCA for the transverse thermal conductivity. The lower bound gives
lowest transverse thermal conductivity as compared to other models.
The effective thermal conductivities of carbon fiber/GNP/epoxy
UHCs are estimated by the micromechanical method developed in Sec­
tion 2. To generate the numerical results following values are used in the
micromechanical modeling: KM = 0.2 W/mK, KFA = 183 W/mK, KFT = 30
W/mK, l = 1.7 μm, t = 60 nm, RK = 3.5 × 10− 9 m2K/W, fGNP = 5 %,
unless otherwise stated. The dispersion of GNPs is assumed to be uni­
form and random inside the epoxy resin.
First, to explore the reinforcement effect of GNPs on the thermal
transport behavior of UHCs, micromechanical analysis is carried out.
The thermal conductivities along the axial and transverse directions as a
functions of carbon fiber volume fraction are shown in Fig. 14 a and b,
respectively. The results of the upper and lower bounds as well as the
CCA approach are included in the figure. The addition of GNPs is
thought to have an insignificant effect on the axial thermal conductivity
of the UHCs as depicted Fig. 14a. It is attributed to the fact that the axial
heat conduction in unidirectional fiber-reinforced composite materials
Fig. 11. Effect of GNP thermal conductivity on the thermal conductivity of the
is dominated by the fiber properties. The thermal conductivity of the
GNP-modified epoxy nanocomposites.
UHCs along the axial direction linearly increases with the increase of
carbon fiber volume fraction. As can be observed in Fig. 14b, the
thickness-dependent variable. It has been reported in the available transverse thermal conductivity displays a significant enhancement with
literature that the thermal conductivity of the graphene may be up to the addition of GNPs. For instance, at 50 vol% carbon fiber, the trans­
5300 W/mK [14,34]. On the other hand, the thermal conductivity of the verse thermal conductivity of the UHCs filled by 5 vol% GNPs increases
GNP remarkably depends on its thickness [33,34]. The value of Kg de­ by 930 % in comparison with the carbon fiber/epoxy composites. The
creases as the GNP thickness increases. The experimental measurements reason is that the transverse heat conduction in unidirectional fiber-
[48] are included in Fig. 11. When the thermal conductivity of GNP is reinforced composites is dominated by the matrix properties. As ex­
derived from Eq. (2), a good agreement is observed between the results pected from the properties of graphene, since a small amount of GNPs
of the micromechanical modeling and experiment [48]. significantly enhances the thermal conductivity of the enriched epoxy
McIvor et al. [50] measured both the axial and transverse matrix therefore, the transverse thermal conductivity of the UHCs can be
temperature-dependent thermal conductivities of carbon increased. The presence of nano-graphene particles in the polymer
fiber-reinforced epoxy composites. As another verification of the model, matrix can serve to promote the formation of effective heat transfer
the calculations are compared with the experimental data [50], as pathways between adjacent micron-sized carbon fibers embedded in the
shown in Fig. 12. The effective thermal conductivities of the carbon hybrid composite system. Thus, the carbon fiber/graphene/polymer
fiber/epoxy composites are plotted as a function of temperature (T). UHCs can transfer heat in higher efficiency to the surrounding envi­
Based on Ref. [50], the temperature-dependent thermal conductivity of ronment compared to the traditional CFRPCs which is a crucial issue for
the constituents is industrial applications. The transverse thermal conductivity of the UHCs
for carbon fiber: KFA = 173.3 + 0.488 T − 0.00178 T2 W/m◦ C and is nonlinearly increased with the increase of carbon fiber volume
KFT = 30 W/m◦ C fraction.
for epoxy: KM = 0.5655 + 0.00124 T W/m◦ C. The variation of carbon fiber/GNP/epoxy UHC thermal conductiv­
where kFA and kFT stand for the axial and transverse thermal conduc­ ities with the GNP volume fraction along the axial and transverse di­
tivities of carbon fiber. The model predictions are seen to be in good rections can be observed in Fig. 15 a and b, respectively. The GNP
agreement with the test results [50]. content varies from 0 to 10 % by volume. The thermal transport per­
To further validate the presented model, the thermal conductivities formance of UHCs in transverse direction is improved by increasing the
of the carbon fiber/epoxy composites predicted by the SUCM are GNPs, as expected from the properties of graphene. The increase of GNP
compared with the predictions of the composite cylinder assemblage content contributes to the increase in GNP distribution density in the
(CCA) approach. Fig. 13 shows this comparison. The longitudinal (KL ) polymer matrix, and more thermal conductive paths can be created to
and transverse (KT ) thermal conductivities of the fiber-reinforced com­ bridge adjacent fibers, which finally results in a higher transverse
posites are estimated by the CCA as [28]. thermal conductivity of the UHCs.
The geometrical characteristics of the GNP on the thermal properties
KL = fF KF + (1 − fF )KM (17)
of carbon fiber/GNP/epoxy UHCs are examined. The thermal conduc­
/ tivities as a function of GNP length are presented in Fig. 16. The UHC
g1 (1 + fF ) + 1 − fF
KT = KM , g 1 = KF KM . (18) thermal conductivities asymptotically increase with increasing the GNP
g1 (1 − fF ) + 1 + fF
length up to l = 10 μm. It is much noticeable in the case of transverse
where fF and KF are the volume fraction and thermal conductivity of the thermal conductivities as shown in Fig. 16b. The improvement in the
fiber. Moreover, the results of the upper and lower bounds are included thermal conductivities may be attributable to the reduced effect of ITR
in Fig. 13. The thermal conductivities of the unidirectional composites between the graphene and polymer matrix which leads to an increase in
can be predicted by the upper and lower bounds [51], respectively, as the thermal conductivity of the enriched epoxy matrix. Furthermore, the
formation of thermal conductive paths between adjacent carbon fibers
Kub = fF KF + (1 − fF )KM (19) becomes easier by the increase of GNP length. At 50 vol% carbon fiber,

8
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 12. Comparison of the model predictions and experimental measurements [50] of the (a) axial and (b) transverse thermal conductivities of the unidirectional
carbon fiber-reinforced epoxy composites.

Fig. 13. Comparison between the results of the SUCM, CCA, upper and lower bounds for the (a) axial and (b) transverse thermal conductivities of the unidirectional
carbon fiber-reinforced epoxy composites.

9
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 14. Effect of GNPs on the (a) axial and (b) transverse thermal conductivities of UHCs.

Fig. 15. Variation of (a) axial and (b) transverse thermal conductivities of UHCs with the GNP volume fraction.

Fig. 16. Variation of (a) axial and (b) transverse thermal conductivities of UHCs with the GNP length.

the value of transverse thermal conductivity increases from 6.2 W/mK performance for the composite structures.
for l = 1.7 μm to 19.98 W/mK for l = 10 μm, i.e., an increase of 222 %. To assess the role of ITR between the GNP and epoxy matrix in the
Accordingly, the ultimate heat transfer ability of UHCs can be hugely thermal transport performance of UHCs, the micromechanical analysis
enhanced by the use of higher lengths of GNPs. Fig. 17 a and b show the is carried out in the absence and the presence of ITR as revealed in
axial and transverse thermal conductivities of the carbon fiber/GNP/ Fig. 18. The high surface area-to-volume ratio of GNPs causes a large
epoxy UHCs, respectively, versus the GNP thickness. The thermal con­ interfacial region between the nano-graphene and matrix. The phonons
ductivity of the hybrid composites, especially in transverse direction, are scattered at the GNP/epoxy interface due to the discrepancy in the
increases as the GNP thickness decreases. It may be due to the increase in phonon density of state. This leads to that the thermal transport is
the GNP thermal conductivity by the decrease of GNP thickness as extremely hindered. The axial thermal conductivity is not influenced by
shown in Eq. (2). In general, the results prove that GNPs with high aspect the ITR according to the results of Fig. 18a. This may be explained by the
ratio allow heat to transfer rapidly and have a better thermal dominant role of carbon fiber properties in axial thermal conductivity of

10
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 17. Variation of (a) axial and (b) transverse thermal conductivities of UHCs with the GNP thickness.

Fig. 18. Effect of ITR on the (a) axial and (b) transverse thermal conductivities of UHCs.

Fig. 19. Variation of (a) axial and (b) transverse thermal conductivities of UHCs with the ITR.

the UHCs. On the basis of the results shown in Fig. 18b, the UHCs in the transfer at the GNP/polymer interfacial region becomes technologically
presence of ITR exhibit lower thermal conductivity in transverse direc­ significant for applications for which the UHCs with very high heat
tion as compared to the UHCs in the absence of ITR. The presence of the dissipation ability are needed.
interface thermal resistance across the GNP/polymer interface leads to a In order to illustrate the role of dispersion of GNPs predicted by our
substantial reduction in the thermal conductivity of the enriched poly­ approach in the UHC thermal properties, a comparative study is per­
mer matrix which finally decreases the transverse thermal conductivity formed. For two conditions of nano-graphene dispersion, including
of the UHCs. Note that matrix properties play the dominant role in uniformly dispersed GNPs, and agglomerated state of GNPs with ξ = 0.5
effective properties of the fibrous composite systems. The relationship and ζ = 0.85, the UHC thermal conductivities are predicted and the
between the thermal conductivities of the UHCs along the axial and numerical results are given in Fig. 20. The UHCs containing uniformly
transverse directions according to the value of GNP/epoxy ITR is shown dispersed GNPs have much higher transverse thermal conductivity than
in Fig. 19 a and b, respectively. In general, a low resistance to heat those containing agglomerated GNPs. From the physical point of view,

11
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

Fig. 20. Effect of GNP agglomeration on the (a) axial and (b) transverse thermal conductivities of UHCs.

Fig. 21. Variation of (a) axial and (b) transverse thermal conductivities of UHCs with parameter ζ

Fig. 22. Effect of GNP alignment on the (a) axial and (b) transverse thermal conductivities of UHCs.

the UHCs containing uniformly dispersed GNPs are much more efficient shows the relation between the effective thermal conductivities and
in constructing thermal conductive paths between adjacent carbon fi­ agglomeration parameters (ξ, ζ). The transverse thermal conductivity of
bers. However, the agglomeration of GNPs in the polymer matrix is not the UHCs is found to rise with increasing ζ up to ξ = ζ and then, it
beneficial to form thermal conductive paths. In other words, graphene sharply decreases. The highest thermal conductivity in the transverse
agglomeration is found to destroy the heat transfer pathways between direction is observed when ξ = ζ; i.e. GNPs are uniformly dispersed in
adjacent carbon fibers. The uniform dispersion of GNPs provides higher the epoxy matrix. Under such condition, the graphene nanoparticles
thermal conductivity enhancement for the epoxy matrix as compared to within the polymer matrix would likely touch each other or link the
the agglomerated state. As a result, producing the hybrid composites adjacent carbon fibers.
with homogeneous dispersion of nanoparticles would offer significant Fig. 22 compares the effects of the GNP alignment on the thermal
advantages over hybrid composites with non-uniform dispersion of transport performance of the carbon fiber/GNP/epoxy UHCs. The nu­
nanoparticles in terms of the thermal transport performance. Fig. 21 merical results for the uniformly (or perfectly) dispersed GNPs in the

12
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

fiber axis is oriented at the thermal loading direction. The dominate role
belongs to the matrix properties when the thermal loading is normal to
the carbon fiber axis.
The effect of carbon fiber arrangement within the GNP-modified
epoxy nanocomposite matrix on the thermal conductivity of UHCs is
investigated, as shown in Fig. 24. The comparison is made with random
and regular arrangements of the carbon fibers. Both square and hexag­
onal packing arrays are reflected as the regular arrangement. Also, the
fiber random arrangement is assumed to be under uniform statistical
distribution. For this case, the RVE is extended to 50 × 50 sub-cells. It is
observed from Fig. 24a that the axial thermal properties are not affected
by different types of fiber arrangement. But, the outcomes of Fig. 24b
reveal that the square array provides higher thermal conductivity
enhancement along the transverse direction as compared to the random
and hexagonal arrays. The random and the hexagonal arrangements
yield almost the same results for the transverse thermal conductivities.
Generally, the axial properties of fibrous composites are not affected by
the fiber arrangement. But, regular arrangement of fiber with square
array can provide better material properties for the fiber-reinforced
composites [52]. Thus, the highly heat-dissipating UHCs can be fabri­
cated by a square array of continuous fibers.
On the basis of available literature, one can expect an anisotropic
conductivity of GNPs [34]. Using the micromechanical model, an
Fig. 23. Variation of thermal conductivity of UHCs with the off-axis angle of analysis is performed to investigate the effect of anisotropy of graphene
carbon fiber. on the thermal conductivity of the GNP-filled epoxy nanocomposites as
shown in Fig. 25. The results indicate that the transverse thermal con­
polymer matrix are included in this figure. Also, the micromechanical ductivity of the graphene has negligibly influence on the nanocomposite
analysis is performed in the presence and absence of ITR. The axial thermal conductivity. However, the axial thermal conductivity of the
thermal conductivity is very negligibly influenced by the variation of GNPs significantly affects the final thermal conductivity of the epoxy
GNP orientation and ITR, as can be seen in Fig. 22a. In contrast, the GNP
orientation significantly contributes to the UHC thermal conductivity in
transverse direction as shown in Fig. 22b. The improvement in the
transverse thermal conductivity of the UHC is maximized when GNPs
are perpendicular to the carbon fibers. The aligned GNPs significantly
enhance the matrix thermal conductivity along the perpendicular to the
carbon fiber.
Fig. 23 displays the predictions of the thermal conductivity of carbon
fiber/GNP/epoxy UHCs as a function of off-axis angle of carbon fibers.
Also, the micromechanical results without GNPs are presented in this
figure. The rate of improvement in thermal conductivity with respect to
the off-axis angle of carbon fibers is increased after the addition of GNPs
within the epoxy resin. In other words, when the fiber orientation varies
from 0◦ to 90◦ , the improvement in the thermal conductivity becomes
more prominent. When the thermal loading is varied from directly in
line with the fiber axis to normal to the fiber axis so-called off-axis
loading, the thermal conductivities of the composite systems vary from
fiber-dominated to matrix-dominated. The material properties of carbon Fig. 25. Effect of GNP anisotropy on the thermal conductivity of the GNP-filled
epoxy nanocomposites.
fiber play the dominant role in the UHC thermal conductivity when the

Fig. 24. Effect of carbon fiber array on the (a) axial and (b) transverse thermal conductivities of UHCs.

13
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

the work reported in this paper.

References

[1] Y. Fu, H. Li, W. Cao, Enhancing the interfacial properties of high-modulus carbon
fiber reinforced polymer matrix composites via electrochemical surface oxidation
and grafting, Compos. Appl. Sci. Manuf. 130 (2020) 105719.
[2] T. Tian, K.D. Cole, Anisotropic thermal conductivity measurement of carbon-fiber/
epoxy composite materials, Int. J. Heat Mass Tran. 55 (23–24) (2012) 6530–6537.
[3] H. Dhieb, J.G. Buijnsters, F. Eddoumy, L. Vázquez, J.P. Celis, Surface and sub-
surface degradation of unidirectional carbon fiber reinforced epoxy composites
under dry and wet reciprocating sliding, Compos. Appl. Sci. Manuf. 55 (2013)
53–62.
[4] S. Aldajah, Y. Haik, Transverse strength enhancement of carbon fiber reinforced
polymer composites by means of magnetically aligned carbon nanotubes, Mater.
Des. 34 (2012) 379–383.
[5] S. Zhang, L. Gao, J. Han, Z. Li, G. Zu, X. Ran, Y. Sun, Through-thickness thermal
conductivity enhancement and tensile response of carbon fiber-reinforced polymer
composites, Compos. B Eng. 165 (2019) 183–192.
[6] Z. Fang, M. Li, S. Wang, Y. Gu, Y. Li, Z. Zhang, Through-thickness thermal
Fig. 26. Axial thermal conductivity of the UHC versus carbon fiber vol­
conductivity enhancement of carbon fiber composite laminate by filler network,
ume fraction. Int. J. Heat Mass Tran. 137 (2019) 1103–1111.
[7] E. Kandare, A.A. Khatibi, S. Yoo, R. Wang, J. Ma, P. Olivier, C.H. Wang, Improving
the through-thickness thermal and electrical conductivity of carbon fibre/epoxy
nanocomposites. laminates by exploiting synergy between graphene and silver nano-inclusions,
Carbon fibers used in structural applications generally can have Compos. Appl. Sci. Manuf. 69 (2015) 72–82.
thermal conductivities about 10 W/mK [53]. By the proposed micro­ [8] M.K. Hassanzadeh-Aghdam, M.J. Mahmoodi, J. Jamali, Effect of CNT coating on
the overall thermal conductivity of unidirectional polymer hybrid nanocomposites,
mechanical method, a sensitive study is performed to investigate the
Int. J. Heat Mass Tran. 124 (2018) 190–200.
effect of carbon fiber thermal conductivity on the final thermal con­ [9] F. Wang, X. Cai, Improvement of mechanical properties and thermal conductivity
ductivities of the UHCs along the axial direction. The micromechanical of carbon fiber laminated composites through depositing graphene nanoplatelets
results are shown in Fig. 26. It is found that adding GNPs in the polymer on fibers, J. Mater. Sci. 54 (5) (2019) 3847–3862.
[10] A. Patti, P. Russo, D. Acierno, S. Acierno, The effect of filler functionalization on
matrix can enhance the axial thermal conductivity of the UHCs as the dispersion and thermal conductivity of polypropylene/multi wall carbon
fiber thermal conductivity is low about 10 W/mK. For example, when nanotubes composites, Compos. B Eng. 94 (2016) 350–359.
the carbon fiber volume fraction is 50 %, the incorporation of GNPs at a [11] H. Kim, A.A. Abdala, C.W. Macosko, Graphene/polymer nanocomposites,
Macromolecules 43 (16) (2010) 6515–6530.
loading of 5 vol% results in approximately 26 % enhancement in the [12] J. Rafiee, M.A. Rafiee, Z.Z. Yu, N. Koratkar, Superhydrophobic to superhydrophilic
axial thermal conductivity of UHCs. According to the results shown in wetting control in graphene films, Adv. Mater. 22 (19) (2010) 2151–2154.
Fig. 14a, addition of GNPs does not affect the axial thermal conductivity [13] H. Fang, S.L. Bai, C.P. Wong, Microstructure engineering of graphene towards
highly thermal conductive composites, Compos. Appl. Sci. Manuf. 112 (2018)
of the UHCs when the carbon fiber thermal conductivity is high. 216–238.
[14] A.A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, C.N. Lau,
4. Conclusions Superior thermal conductivity of single-layer graphene, Nano Lett. 8 (3) (2008)
902–907.
[15] Z. Xu, S. He, J. Zhang, S. Huang, A. Chen, X. Fu, P. Zhang, Relationship between
A hierarchical micromechanics technique was developed to analyze the structure and thermal properties of polypropylene/graphene nanoplatelets
the effect of adding the GNPs within the polymer matrix on the thermal composites for different platelet-sizes, Compos. Sci. Technol. 183 (2019) 107826.
[16] Y.H. Zhao, Y.F. Zhang, S.L. Bai, X.W. Yuan, Carbon fibre/graphene foam/polymer
transport performance of the carbon fiber/GNP/epoxy UHCs. The GNPs composites with enhanced mechanical and thermal properties, Compos. B Eng. 94
improved the transverse thermal conductivity of the UHCs. A high (2016) 102–108.
content, high length, and low thickness of the GNPs led to the [17] M. Owais, J. Zhao, A. Imani, G. Wang, H. Zhang, Z. Zhang, Synergetic effect of
hybrid fillers of boron nitride, graphene nanoplatelets, and short carbon fibers for
improvement in the thermal conductivity along the transverse direction. enhanced thermal conductivity and electrical resistivity of epoxy nanocomposites,
Adding the GNPs in the polymer matrix enhanced the axial thermal Compos. Appl. Sci. Manuf. 117 (2019) 11–22.
conductivity of the UHCs when the fiber thermal conductivity is low. [18] B. Zhang, R. Asmatulu, S.A. Soltani, L.N. Le, S.S. Kumar, Mechanical and thermal
properties of hierarchical composites enhanced by pristine graphene and graphene
The agglomeration of GNPs within the epoxy matrix decreased the ca­
oxide nanoinclusions, J. Appl. Polym. Sci. 131 (19) (2014).
pacity of heat dissipation. Besides, the ITR between the GNP and epoxy [19] L. Liu, L. Xiao, X. Zhang, M. Li, Y. Chang, L. Shang, Y. Ao, Improvement of the
matrix led to the reduction of thermal transport properties. A uniform thermal conductivity and friction performance of poly (ether ether ketone)/carbon
dispersion of GNPs and low ITR were effective routs to obtain the fiber laminates by addition of graphene, RSC Adv. 5 (71) (2015) 57853–57859.
[20] Y.J. Noh, S.Y. Kim, Synergistic improvement of thermal conductivity in polymer
thermally conductive UHCs with the higher performance. It was composites filled with pitch based carbon fiber and graphene nanoplatelets, Polym.
observed that GNPs can be perpendicular to the carbon fibers with the Test. 45 (2015) 132–138.
objective of further enhancing the transverse thermal conductivity. The [21] J. Yu, T.E. Lacy Jr., H. Toghiani, C.U. Pittman Jr., Micromechanically-based
effective thermal conductivity estimates for polymer nanocomposites, Compos. B
influences of several critical microstructure aspects related to the carbon Eng. 53 (2013) 267–273.
fibers, including the volume fraction, orientation and arrangement type [22] S.Y. Kim, J.U. Jang, B.F. Haile, M.W. Lee, B. Yang, Swarm intelligence integrated
on the UHC thermal transport properties were examined. Heat can be micromechanical model to investigate thermal conductivity of multi-walled carbon
nanotube-embedded cyclic butylene terephthalate thermoplastic nanocomposites,
more efficiently dissipated by the square array of carbon fibers Compos. Appl. Sci. Manuf. 128 (2020) 105646.
embedded in the GNP-epoxy matrix thereby reducing thermal damage. [23] N. Bonfoh, C. Dreistadt, H. Sabar, Micromechanical modeling of the anisotropic
According to some comparisons, there was a good agreement between thermal conductivity of ellipsoidal inclusion-reinforced composite materials with
weakly conducting interfaces, Int. J. Heat Mass Tran. 108 (2017) 1727–1739.
the model predictions and experimental measurements. The developed [24] K. Chu, C.C. Jia, W.S. Li, Effective thermal conductivity of graphene-based
micromechanics technique gives a more accurate insight into the ther­ composites, Appl. Phys. Lett. 101 (12) (2012) 121916.
mal transport mechanisms of the general UHC systems which can be [25] M. Pakseresht, R. Ansari, M.K. Hassanzadeh-Aghdam, Laminate analogy approach
for the effective elastic properties of metal matrix nanocomposites filled with
utilized as structural materials in the next-generation.
randomly dispersed graphene nanoplatelets, Proc. IME C J. Mech. Eng. Sci. 234 (6)
(2020) 1212–1219.
[26] K. Chu, W.S. Li, H. Dong, Role of graphene waviness on the thermal conductivity of
Declaration of competing interest graphene composites, Appl. Phys. A 111 (1) (2013) 221–225.
[27] C.W. Nan, G. Liu, Y. Lin, M. Li, Interface effect on thermal conductivity of carbon
The authors declare that they have no known competing financial nanotube composites, Appl. Phys. Lett. 85 (16) (2004) 3549–3551.
interests or personal relationships that could have appeared to influence

14
M.K. Hassanzadeh-Aghdam et al. International Journal of Thermal Sciences 171 (2022) 107209

[28] S.I. Kundalwal, M.C. Ray, Estimation of thermal conductivities of a novel fuzzy nanocomposite cylinders by a mesh-free method, Acta Mech. 224 (11) (2013)
fiber reinforced composite, Int. J. Therm. Sci. 76 (2014) 90–100. 2817–2832.
[29] S.T. Huxtable, D.G. Cahill, S. Shenogin, L. Xue, R. Ozisik, P. Barone, P. Keblinski, [42] Q.S. Yang, X.Q. He, X. Liu, F.F. Leng, Y.W. Mai, The effective properties and local
Interfacial heat flow in carbon nanotube suspensions, Nat. Mater. 2 (11) (2003) aggregation effect of CNT/SMP composites, Compos. B Eng. 43 (1) (2012) 33–38.
731–734. [43] S.K. Mital, P.L. Murthy, R.K. Goldberg, Micromechanics for particulate-reinforced
[30] M.K. Hassanzadeh-Aghdam, M.J. Mahmoodi, Micromechanical modeling of composites, Mech. Compos. Mater. Struct. Int. J. 4 (3) (1997) 251–266.
thermal conducting behavior of general carbon nanotube-polymer [44] M.K. Hassanzadeh-Aghdam, M.J. Mahmoodi, R. Ansari, Micromechanics-based
nanocomposites, Mater. Sci. Eng., B 229 (2018) 173–183. characterization of mechanical properties of fuzzy fiber-reinforced composites
[31] H.S. Kim, J.U. Jang, J. Yu, S.Y. Kim, Thermal conductivity of polymer composites containing carbon nanotubes, Mech. Mater. 118 (2018) 31–43.
based on the length of multi-walled carbon nanotubes, Compos. B Eng. 79 (2015) [45] M. Haghgoo, M.K. Hassanzadeh-Aghdam, R. Ansari, Effect of piezoelectric
505–512. interphase on the effective magneto-electro-elastic properties of three-phase smart
[32] L. Hu, T. Desai, P. Keblinski, Thermal transport in graphene-based nanocomposite, composites: a micromechanical study, Mech. Adv. Mater. Struct. 26 (23) (2019)
J. Appl. Phys. 110 (3) (2011), 033517. 1935–1950.
[33] W.R. Zhong, M.P. Zhang, B.Q. Ai, D.Q. Zheng, Chirality and thickness-dependent [46] A. Sayyidmousavi, H. Bougherara, S.R. Falahatgar, Z. Fawaz, Prediction of the
thermal conductivity of few-layer graphene: a molecular dynamics study, Appl. effective thermal conductivity of fiber reinforced composites using a
Phys. Lett. 98 (11) (2011) 113107. micromechanical approach, J. Mech. 35 (2) (2019) 179–185.
[34] K. Chu, W.S. Li, F.L. Tang, Flatness-dependent thermal conductivity of graphene- [47] S.I. Kundalwal, R.S. Kumar, M.C. Ray, Effective thermal conductivities of a novel
based composites, Phys. Lett. 377 (12) (2013) 910–914. fuzzy carbon fiber heat exchanger containing wavy carbon nanotubes, Int. J. Heat
[35] F. Deng, Q.S. Zheng, L.F. Wang, C.W. Nan, Effects of anisotropy, aspect ratio, and Mass Tran. 72 (2014) 440–451.
nonstraightness of carbon nanotubes on thermal conductivity of carbon nanotube [48] A. Yu, P. Ramesh, M.E. Itkis, E. Bekyarova, R.C. Haddon, Graphite nanoplatelet-
composites, Appl. Phys. Lett. 90 (2) (2007), 021914. epoxy composite thermal interface materials, J. Phys. Chem. C 111 (21) (2007)
[36] L.C. Tang, Y.J. Wan, D. Yan, Y.B. Pei, L. Zhao, Y.B. Li, G.Q. Lai, The effect of 7565–7569.
graphene dispersion on the mechanical properties of graphene/epoxy composites, [49] K.M. Shahil, A.A. Balandin, Thermal properties of graphene and multilayer
Carbon 60 (2013) 16–27. graphene: applications in thermal interface materials, Solid State Commun. 152
[37] Z. Shokrieh, M.M. Shokrieh, Z. Zhao, A modified micromechanical model to predict (15) (2012) 1331–1340.
the creep modulus of polymeric nanocomposites, Polym. Test. 65 (2018) 414–419. [50] S.D. McIvor, M.I. Darby, G.H. Wostenholm, B. Yates, L. Banfield, R. King, A. Webb,
[38] Z. Li, J. Chu, C. Yang, S. Hao, M.A. Bissett, I.A. Kinloch, R.J. Young, Effect of Thermal conductivity measurements of some glass fibre-and carbon fibre-
functional groups on the agglomeration of graphene in nanocomposites, Compos. reinforced plastics, J. Mater. Sci. 25 (7) (1990) 3127–3132.
Sci. Technol. 163 (2018) 116–122. [51] D. Kumlutas, I.H. Tavman, A numerical and experimental study on thermal
[39] R. Rafiee, A. Eskandariyun, Predicting Young’s modulus of agglomerated conductivity of particle filled polymer composites, J. Thermoplast. Compos. Mater.
graphene/polymer using multi-scale modeling, Compos. Struct. (2020) 112324. 19 (4) (2006) 441–455.
[40] D.L. Shi, X.Q. Feng, Y.Y. Huang, K.C. Hwang, H. Gao, The effect of nanotube [52] M.M. Aghdam, A. Dezhsetan, Micromechanics based analysis of randomly
waviness and agglomeration on the elastic property of carbon nanotube-reinforced distributed fiber reinforced composites using simplified unit cell model, Compos.
composites, J. Eng. Mater. Technol. 126 (3) (2004) 250–257. Struct. 71 (3–4) (2005) 327–332.
[41] R. Moradi-Dastjerdi, A. Pourasghar, M. Foroutan, The effects of carbon nanotube [53] K. Mishra, B. Garnier, S. Le Corre, N. Boyard, Accurate measurement of the
orientation and aggregation on vibrational behavior of functionally graded longitudinal thermal conductivity and volumetric heat capacity of single carbon
fibers with the 3 ω method, J. Therm. Anal. Calorim. 139 (2) (2020) 1037–1047.

15

You might also like