You are on page 1of 22

Materials Science & Engineering R 132 (2018) 1–22

Contents lists available at ScienceDirect

Materials Science & Engineering R


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

Thermal conductivity of polymers and polymer nanocomposites T


a,b b b,c,⁎
Congliang Huang , Xin Qian , Ronggui Yang
a
School of Electrical and Power Engineering, China University of Mining and Technology, Xuzhou, 221116, PR China
b
Department of Mechanical Engineering, University of Colorado, Boulder, Colorado, 80309, USA
c
Materials Science and Engineering Program, University of Colorado, Boulder, Colorado, 80309, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Polymers are widely used in industry and in our daily life because of their diverse functionality, light weight, low
Thermal conductivity cost and excellent chemical stability. However, on some applications such as heat exchangers and electronic
Polymer packaging, the low thermal conductivity of polymers is one of the major technological barriers. Enhancing the
Nanocomposite thermal conductivity of polymers is important for these applications and has become a very active research topic
Polymer chain
over the past two decades. In this review article, we aim to: 1). systematically summarize the molecular level
Network
Interface
understanding on the thermal transport mechanisms in polymers in terms of polymer morphology, chain
structure and inter-chain coupling; 2). highlight the rationales in the recent efforts in enhancing the thermal
conductivity of nanostructured polymers and polymer nanocomposites. Finally, we outline the main advances,
challenges and outlooks for highly thermal-conductive polymer and polymer nanocomposites.

1. Introduction chain coupling; 2). highlight the rationales in the recent efforts in en-
hancing the thermal conductivity of nanostructured polymers and
Polymers and polymer composites are used ubiquitously in a wide polymer nanocomposites.
range of industrial applications ranging from structural materials to The thermal conductivity of bulk polymers is usually very low, on
electronics and in our daily life from chopsticks to trash bins due to the order of 0.1–0.5 W·m−1 K−1, which is due to the complex mor-
their diverse functionality, light weight, low cost, and excellent che- phology of polymer chains [6]. Fig. 1(a) shows a typical structure of a
mical stability. However, the low thermal conductivity of polymers polymer, which consists of crystalline domains where polymer chains
limits in their applications in some fields. For example, the low thermal are aligned periodically, and amorphous domains where the polymer
conductivity of polymers can be one of the major technological barriers chains are randomly entangled. The thermal conductivity of a polymer
for the polymer-based flexible electronics due to the limited heat depends greatly on its morphology. When amorphous domains are
spreading capability [1–3]. If a polymer can be engineered with high dominant, vibrational modes in the polymer tend to be localized, re-
thermal conductivity, polymeric heat spreaders and heat exchanger can sulting in a low thermal conductivity. It is therefore natural to expect
be manufactured with superb features including structural compact- that thermal conductivity can be enhanced by improving the alignment
ness, light weight, resistance against corrosions, and ease of processing of polymer chains. Indeed many efforts have been devoted to align
and low-cost, which could in turn find many applications in electronics, polymer chains to enhance the thermal conductivity, by using me-
water and energy industry. [4,5] Thus, enhancing the thermal con- chanical stretching, nanoscale templating and electrospinning. A
ductivity of polymers and polymer composites are of great interests. thermal conductivity as high as 104 W·m−1 K−1 has been achieved for
Over the past two decades, with a better understanding of the polyethylene (PE) after stretching with a draw ratio of 400 for nano-
fundamental heat transfer process at the micro-, nano- and even mo- fibers with a diameter of 50–500 nm and lengths up to tens of milli-
lecular- scales, there have been significant efforts devoting to enhan- meters [7]. The thermal conductivity was shown to be enhanced for
cing the thermal conductivity of polymers and polymer nanocompo- more than 20 times in polythiophene nanofibers with a fiber diameter
sites, which are expected to enable a broader range of applications. In of about 50 nm, prepared using templated electropolymerization [8].
this review, we aim to: 1). systematically summarize the understanding Inspired by these experimental efforts, molecular dynamics simulations
on the physical mechanisms that controls the thermal transport in have been conducted to understand how nanoscale structures affect the
polymers by relating those to polymer chain morphology and inter- thermal conductivity. As shown in Fig. 1(b), in addition to polymer


Corresponding author at: Department of Mechanical Engineering, University of Colorado, Boulder, Colorado, 80309, USA.
E-mail address: Ronggui.Yang@Colorado.Edu (R. Yang).

https://doi.org/10.1016/j.mser.2018.06.002
Received 15 May 2018; Accepted 25 June 2018
Available online 09 July 2018
0927-796X/ © 2018 Elsevier B.V. All rights reserved.
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Nomenclature PA6 polyamide-6


PBX polymer-bonded explosives
0D 0- dimensional PC polycarbonate
1D 1- dimensional PDA polydopamine
2D 2- dimensional PE polyethylene
3D three dimensional PEEK poly(ether-ether-ketone)
Al2O3 aluminum oxide PEG Poly(ethylene glycol)
AlN aluminum nitride P-GC phenyl-aminated GO and CNT hybrid fillers
BB Bottlebrush PHB poly(3-hydroxylbutyrate)
BN boron nitride PI polyimide
BNNS boron-nitride nanosheets PMMA poly(methyl methacrylate)
CE cyanate ester PP Polypropylene
CF carbon foam PPS polyphenylene sulfide
CNT carbon nanotubes PS polystyrene
EG expanded graphite PVS Poly(vinyl alcohol)
E-G Cethyl-aminated graphene-oxide and CNT hybrid fillers PVDF poly(vinylidene fluoride)
EMT effective medium theory PVP Polyvinylpyrrolidone
GF graphene foam PVPh poly(4-vinyl phenol)
GNP graphene nanoplate R-GC raw GO and CNT hybrid fillers
GNR graphene nanoribbons Si3N4 silicon nitride
GO graphene oxide SiC silicon carbide
GS graphene sheet SR silicone rubber
H-bond Hydrogen bond SWCNT Single-wall CNT
h-BN hexagonal boron nitride UHMW- ultrahigh molecular weight
HDPE high density polyethylene PE polyethylene
MD Molecular dynamics vdW van der Waal
MgO magnesium oxide ZnO zinc oxide
MWCNT Multi-wall CNT P(VDF- poly(vinylidene
P3HT poly(3-hexylthiophene-2,5-diyl) TeFE fluoride-trifluoroethylene)
PAA Poly(acrylic acid)

Fig. 1. Schematic diagrams of a polymer: (a) the morphology of a polymer consisting of crystalline and amorphous domains; (b) structure of a polymer chain.

chain alignment, the thermal conductivity of a polymer also depends on interfacial thermal resistance, and the concentration and the geometric
the structure of chains including backbone bonds and side chains, and shapes of fillers. When the filler concentration is large enough, high
the inter-chain coupling. conductivity fillers might form thermally conductive networks, as
In addition to engineering the morphology of polymer chains, an- shown in Fig. 2(b). Although nanocomposites with filler network could
other common method to enhance the thermal conductivity of polymers possess a higher thermal conductivity than that without a network,
is to blend polymers with highly thermal conductive fillers. The pro- their thermal conductivity could still be low due to the large inter-filler
gress of nanotechnology over the last two decades not only provides thermal contact resistance. Recently, three-dimensional fillers, such as
more diverse high thermal conductivity fillers of different material carbon and graphene foams, have drawn a lot of attention. The fun-
types and topological shapes but also advances the understanding at the damental thermal transport mechanisms and recent synthesis efforts in
nanoscale. Fig. 2 shows a sketch of a polymer nanocomposite to illus- both types of nanocomposites are reviewed.
trate the thermal transport mechanisms. In general, there are two types This review article is organized as follows. In Section 2, we in-
of polymer nanocomposites depending on whether nano-fillers form a troduce the experimental progress on the enhancement of thermal
network or not. When the filler concentration is low, no inter-filler conductivity by aligning polymer chains, and then review the methods
networks could be formed, as shown in Fig. 2 (a). The thermal con- to further tune the thermal conductivity by engineering chain structure
ductivity is essentially determined by the filler-matrix coupling, i.e, and inter-chain coupling, as illustrated in Fig. 3 (a). In Section 3, we

2
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 2. Schematic diagrams of polymer nanocomposites: (a) without inter-filler network; (b) with inter-filler networks. Thermally conductive pathway is identified
with dash lines.

discuss the thermal conductivity of polymer nanocomposites both with expected to enhance the thermal conductivity of polymers. In crystal-
and without inter-filler networks, as illustrated in Fig. 3(b). line polyethylene (PE) nanofibers, the measured thermal conductivity
could be as high as 104 W·m−1 K-1 [7]. There have already been some
2. Thermal conductivity of pristine polymers experimental methods demonstrated to improve the degree of the chain
alignment. For example, the chain alignment could be enhanced by
In this Section, we discuss the dominant factors controlling the thermal annealing [9–11]. Experimental studies on the poly(vinylidene
thermal transport in pristine polymers and the methods for manip- fluoride-trifluoroethylene) [P(VDF-TrFE)] polymer showed that when
ulating the thermal conductivity through morphology control. One of the polymer film was annealed above its melting temperature (421 K), a
the most intuitive methods for improving the thermal conductivity is to 300–400 % increase of thermal conductivity occurs because of the
improve the order of chain alignment (chain morphology), which has chain alignment [12]. It is also unveiled that the geometric constraints
been experimentally demonstrated. This method is discussed in Section at the substrate-P(VDFTrFE) interface have an important role on the
2.1. Apart from the chain morphology, the chain structure also plays an formation of aligned chains in the process of melt-recrystallization [12],
important role in determining the thermal conductivity. Methods for which has also confirmed by other works [13,14].
manipulating thermal transport by engineering the chain structure in- The mechanisms of how annealing and geometric constraints of the
cluding backbone and side chains are discussed in Section 2.2. In Sec- substrate affect chain alignment of polymers is still not well understood
tion 2.3, we further discuss the methods by enhancing inter-chain and thus it is difficult to further improve the chain alignment. In this
coupling through introducing strong interactions like hydrogen bonds Section, some experimental methods which have been successfully
(H-bonds) and cross-linkers formed by covalent bonds to tune the applied to enhance the chain alignment, such as the mechanical
thermal conductivity of polymers, as compared to the weak van der stretching, nanoscale templating and electrospinning, are summarized
Waals (vdW) in most of polymers. in Sections 2.1.1–2.1.3.

2.1. Alignment of chain orientations 2.1.1. Mechanical stretching


Mechanical stretching could significantly increase the thermal
A typical semi-crystalline polymer contains crystalline domains of conductivity of polymers due to the increased order of chain orienta-
aligned chais, and the amorphous domains with randomly twisted and tion. The first demonstration was performed by Choy et al. [15–17],
entangled chains. The lack of periodicity in the amorphous domains who found that the thermal conductivity of the ultrahigh molecular
severely localizes the vibrational modes, thereby suppressing the weight polyethylene (UHMW-PE) could exceed 40 W·m−1 K−1 when
thermal conductivity. Increasing the chain alignment is therefore the drawing ratio reaches beyond 300 [17], as shown in Fig. 4(a). Using

Fig. 3. Physical mechanisms affecting the thermal conductivity of: (a) polymers; (b) polymer nanocomposites.

3
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 4. Thermal conductivities under mechanical stretching: (a) effect of draw ratios in UHMW-PE [7]. Copyright 2010 Nature Publishing Group; (b) influence of
crystalline in UHMW-PE [20]. Copyright 2017 American Chemical Society; (c) Tensile effects on thermal transport in carbine and cumulene [23]. copyright 2015
Nature Publishing Group.

a two-stage heating method, Shen et al. [7] obtained a higher drawing transport in cumulene is relatively independent of the strains, as shown
ratio (400) and produced a PE nanofiber with a diameter of 50–500 nm in Fig. 4(c). This can be explained by the two competing factors during
where a higher thermal conductivity of 104 W·m−1 K−1 was obtained, stretching, one is the increased phonon lifetime due to the increased
as illustrated in Fig. 4(a). The authors attributed the enhancement of order of chain morphology and the other is decreased group velocity
the thermal conductivity to the improved fiber crystallinity. Some other due to the strains.
experimental works also confirmed the enhancement thermal con-
ductivity due to the enhanced chain orientation through mechanical 2.1.2. Nanoscale templating
stretching [18,19]. Nanoscale templating is another way to align polymer chains.
However, a higher crystallinity does not always lead to a higher Polymers are melted and infiltrated in a porous template such as porous
thermal conductivity, because the thermal conductivity depends on the anodic alumina. Removing the porous anodic alumina template using
overall chain alignment in the polymer rather than the portion of the NaOH aqueous solution left with an array of polymer nanofibers. The
extremely aligned chains (crystalline). An example supporting this alignment of the polymer chains is improved due to the flow of polymer
understanding is shown in Fig. 4(b). The thermal conductivity of a melt in the nanoporous template, and the thermal conductivity can
thermally stretched UHMW-PE microfiber approaches 51 W·m−1 K−1, therefore be enhanced along the axial direction of a fiber.
while its crystallinity is reduced from 92% to 83% during the stretching In the work by Singh et al. [8], aligned arrays of polythiophene
[20]. Such increased thermal conductivity is due to the improved chain nanofibers were electro-polymerized vertically inside the nano-chan-
alignment in the amorphous domains. nels of anodic alumina templates, as shown in Fig. 5(a). The length of
To theoretically understand the enhanced thermal conductivity by the polythiophene fibers is tuned by the charge passing through the
mechanical stretching, there have been quite some research works electrochemical cell and the diameter is controlled by the diameter of
using atomistic simulations. For example, using MD simulations, Liu the pores in the template. The thermal conductivity of polythiophene
and Yang [21] showed that the thermal conductivity of PE increases nanofibers is therefore improved to be 4.4 W·m−1 K−1, more than 20
when the polymer is stretched slowly. Such thermal conductivity en- times higher than that of the bulk polymer, because of improvement of
hancement is found to be strongly correlated with the orientation order the chain alignment as schematically shown in Fig. 5(b).
parameter [22] which describes the change of chain conformation. The thermal conductivity could be increased by decreasing the fiber
However, the tensile extension not always leads to an enhanced thermal diameters with a nanoscale templating method, because of the de-
conductivity. For example, Wang and Lin [23] found that the thermal creased entangle possibility of the chains. The thermal conductivity of

Fig. 5. Nanoscale templating method: (a) schematic diagram of experimental setup for template guided electrochemical synthesis of nanofibers; (b) chain orientation
comparison between bulk polymer (left) and electropolymerized nanofiber (right) [8]. Copyright 2014 Nature Publishing Group.

4
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

approximately 7 W·m−1 K−1 was achieved for both the 200-nm-dia- non-bonding force is usually much smaller than the covalent force and
meter high-density PE (HDPE) fibers and the 100-nm-diameter poly(3- thus is not as important as covalent bonds in affecting the thermal
hexylthiophene-2,5-diyl) (P3HT) fibers [24]. The thermal conductivity conductivity of a continuous single chain, we therefore focus our dis-
of the 200-nm-diameter HDPE nanofiber can be further improved to cussions on the conformation energy associated with covalent bonds. As
26.5 W·m−1 K−1 using the nanoporous template wetting technique. The shown in Fig. 8, the covalent bonding energy can be decomposed into
authors attribute this large increase of thermal conductivity to the in- three parts according to different vibrational modes. The first part is the
tegrative effects of shear rate, vibrational perturbation, translocation, two-body stretching energy Ebond , due to the stretching of a bond
nano-confinement and crystallization. connecting two atoms. In addition to the two-body bond stretching,
many body interactions also contribute to the conformation energy, and
2.1.3. Electrospinning usually up to four body interactions are considered, which are the
Electrospinning is a nanofiber production technique which uses three-body angular bending motion, Eangle , and the four-body dihedral
electrostatic forces to draw charged threads of polymeric solution. A bending motion, Edihedral − angle . The angular bending involves three
polymer is dissolved in a solution and held in a reservoir. Usually a atoms connected by two neighboring covalent bonds, and the three
syringe with a sharp tip is used for dispensing the polymer. By applying atoms have a scissoring like motion that changes the angle between the
an electrostatic force between the tip and a grounded collector plate, two covalent bonds. A dihedral bending motion involves four atoms
polymer solution is drawn to fibers with nano-scale diameters [25]. connected by three nose-to-tail covalent bonds. The dihedral angle is
There are two experimental parameters affecting the chain morphology the angle between the two planes sharing a covalent bond in the middle
of electrospun polymers. The first parameter is the strength of the and the atom at two ends of the atomic quadruplet. The Edihedral − angle
electric field during the electrospinning process. It is reported that describes the energy barrier to overcome if a polymer segment is ro-
strengthening the electric field is beneficial to enhance the chain tated. A larger bond-stretching energy Ebond or/and angular bending
alignment, and thus the thermal conductivity of the electrospun PE energy Eangle means a stiffer backbone, and a larger dihedral-angle en-
nanofibers [26,27]. According to the degree of polymer chain align- ergy Edihedral − angle tends to suppress the segmental rotation. In the fol-
ment in an electrospun fiber, Canetta et al. [28] summarized the pos- lowing, the effect of the backbone stiffness (bond-stretching and angle-
sible orientations of the chains in a polymer as shown in Fig. 6. To bending strengths) and the dihedral-angle-torsion stiffness on thermal
enhance the thermal conductivity, it is preferable that the chains are transport in polymer chains are discussed.
mainly aligned along the axis of the fiber as shown in Fig. 6(a). The Generally, higher stiffness of bond stretching and angular bending
other parameter affecting the alignment of polymer chains is the jet leads to a higher thermal conductivity. One well-known example is PE,
speed. As the evaporation is faster on the outside of the jet in electro- which has very high intrinsic thermal conductivity due to the strong
spinning, a core-shell morphology as shown in Fig. 6(b) could be bonding, where both the bond stretching and bond bending are very
formed in electrospun fibers [29], which is helpful for enhancing the stiff [32,33]. Fig. 9 shows a more detailed thermal conductivity de-
thermal conductivity. Sometimes, the super-molecular morphology as pendence on the stiffness of the polymer chains [34]. Generally, the
shown in Fig. 6(c) could be also formed to enhance the elastic modulus existence of double eC ] Ce bond with sp2 hybridization results in
of electrospun fibers [30], which is usually related to a higher thermal higher thermal conductivity due to two reasons: (1) The bonding en-
conductivity. ergy of eC ] Ce bond is 2.8 times that of a eCeCe with sp3 hy-
Fig. 7 shows the enhanced thermal conductivity of polymers with bridization; (2) Consecutive sp2 bond forms delocalized conjugated π-
chain alignment through different processing techniques. The thermal bond, which constrains the atoms in the conjugated π-bond to be in the
conductivity of the electrospinning PS nanofibers by Canetta et al. [28] same atomic plane. As a result, the conjugated π-bond greatly increases
is found to be between 6.6 and 14.4 W·m−1 K−1, a significant increase the stiffness of segmental rotation. One example is shown in Fig. 9(a),
above the typical thermal conductivity (∼0.15 W·m−1 K−1) of bulk PS where the thermal conductivity of π-conjugated polyacetylene is higher
due to the preferential alignment of molecular chains, along with the than that of PE [35]. A recent work by Xu et al. [36] found that the
reduction in defects and voids compared to bulk. Zhong et al. [31] existence of conjugated eC ] Ce in poly(3-hexylthiophene) (P3HT)
reported enhanced thermal conductivity (1–2 W·m−1 K−1) of Nylon-11
nanofibers fabricated by electrospinning and post-stretching. They re-
vealed that the crystalline morphology plays an important role to en-
hance the thermal conductivity in addition to the chain alignment. In
general, Fig. 7 shows that the thermal conductivity increases with the
decreasing nanofiber diameters, because the fiber surface could limit
the random orientation of polymer chains [28].

2.2. Atomic structure of polymer chains

The structure of polymer chains could also play an important role in


the thermal conductivity. The bond stretching and the angular bending
strengths of the backbone influence the chain entanglement, while the
strength of dihedral bending can also greatly affect the rotation of chain
segments. Both the chain entanglement and the rotation of chain seg-
ments can in turn affect the thermal conductivity while the side-chain
branching out of the backbone has a role as well.

2.2.1. Backbone
In Fig. 8, we use a PE polymer chain as an example to illustrate the
backbone structure and the conformation energy of the chain con- Fig. 6. Orientations of polymer chains: (a) a nanofiber with preferable aligned
tributed by different vibrational modes. The total conformation energy chains; (b) a core-shell morphology with the shell formed by aligned chains and
of a polymer chain can be separated into the conformation energy of the core comprised of randomly oriented chains; (c) a super-molecular morphology
covalent bonds and the non-bonding energy due to van der Waals with aligned-chain grains filled in a randomly oriented chain matrix [28].
(vdW) interactions, Coulomb interactions and hydrogen bonds. The Copyright 2014 American Institute of Physics.

5
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

can simultaneously achieve efficient phonon transport along the chains


and strong noncovalent inter-chain interaction due to the π- π stacking.
As a result, thermal conductivity as large as 2.2 W·m−1 K−1 is measured
in oxidative chemical vapor deposition (oCVD) grown thin films. This
work shows that polymers with conjugated bonds are promising can-
didates for achieving higher thermal conductivity. Conjugated π-bond
is also found in aromatic rings as shown in Fig. 9(b), where the thermal
conductivity of chains with aromatic-backbone structures is much
higher than that with aliphatic-backbone structures. In addition, the
type of carbon-carbon bonding, the thermal conductivity of polymer
chains can also be tuned by replacing the hydrocarbon functional
groups in the backbone with other atomic species. Fig. 9(c) shows the
effect of replacing eCH2- groups in the backbone with O atoms. Be-
cause the eCH2eOe bonding energy (∼335 kJ/mol) is lower than
eCH2eCH2- (∼ 350 kJ/mol) and O atom is heavier than eCH2- group,
the thermal conductivity of poly(ethylene oxide) is much lower than
that of both PE and poly(methylene oxide).
Fig. 7. Measured thermal conductivity for nanofibers of PE [7], Nylon-11 [31], The effect of the dihedral-angle stiffness on the thermal conductivity
amorphous PS (a-PS) [28], and amorphous polythiophene (a-PT) [8] [28].
has also been studied using MD simulations performed in a model
Copyright 2014 American Institute of Physics.
system of bulk PE amorphous polymer [37]. The energy constant (K1) of
the dihedral angle energy which stands for the dihedral-angle stiffness
is systematically changed, and the resultant structures and thermal
conductivity are shown in Fig. 10. When K1 is increased, the radius of
gyration increases, which means more extended chains (Fig. 10a), and
the persistence length also increases, which suggests more straight
chains (Fig. 10a). Because of more extended and straight chains, the
thermal conductivity in all three orthogonal directions (X, Y and Z)
increase with increasing K1, as shown in Fig. 10(b). Further thermal
conductivity decomposition analysis reveals that thermal transport
through covalent bonds dominates the thermal conductivity over other
contributions from the non-bonding vdW interactions and the transla-
tion of molecules (Fig. 10c). The contribution of the non-bonded vdW
force to the thermal conductivity is also enhanced by the increase of
energy constant (K1) of the dihedral angle energy, because of the
shortened inter-atomic distance.

2.2.2. Side-chains
Side-chains are functional groups branching out of the backbones
[38]. Since side-chains change the topology of the polymer chains, the
vibrational dynamics and the thermal conductivity are both expected to
change when different types of side chains are introduced. Here we
discuss the effect of side-chains on the thermal conductivity of poly-
mers.
Fig. 8. Effect of backbone structure of a polymer chain on the thermal con- Thermal conductivity of a polymer chain is found to decrease when
ductivity, illustrated using a PE polymer chain as an example. side-chains are introduced, and usually a heavier side-chain leads to a
lower thermal conductivity one [38], as shown in Fig. 11. It also shows
that the thermal conductivity is decreased when the distance between
two neighboring side-chains is decreased from 100 segments, to 75 and
then to 50 segments. The decreasing trend of the thermal conductivity

Fig. 9. The thermal conductivity dependence of bond strength: (a) effect of double bonds; (b) effect of aromatic backbone; (c) effect of bond-strength disorder [34].
Copyright 2012 American Physical Society.

6
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 10. Effects of the dihedral-angle stiffness: (a) persistence length and radius of gyration of chains; (b) thermal conductivity; (c) contributions of covalent bonding
and non-covalent bonding on the thermal conductivities along X direction [37]. Copyright 2016 American Chemical Society.

insensitive to the side-chain length when this length is larger than 10


segments, as that shown in Fig. 12(b). The mechanism of the reduced
thermal conductivity caused by the side-chains could be understood as
a result of phonon localization and phonon scatterings [38,39]. In bulk
polymers, side chains are also found to decrease the thermal con-
ductivity, because the crystallinity are found to decrease dramatically
when side chains are introduced [39].
Interestingly, side-chains could also lead to an increase in thermal
conductivity of polymers. Using the ultrafast laser based pump-probe
technique, Guo et al. [40] showed that the measured thermal con-
ductivity of conjugated polymers with linear and long side-chains in-
crease by 160% compared to that with short side-chains. MD simula-
tions further revealed that these linear and long side-chains tend to
increase the structural order of the conjugated polymers, which results
in an increased thermal conductivity. Chen et al. [41,42] also reported
that the thermal conductivity of polyaniline can be enhanced by in-
troducing functional groups due to the improved chain alignment. Side-
chains can change the thermal conductivity in both ways which calls for
more theoretical works to further understand the effect of side-chains
Fig. 11. Thermal conductivities of PE chains branched with different side-
on the thermal conductivity of polymers.
chains. The side-chains are attached to backbone for every 50, 75, and 100
segments. The black, red and blue columns stand for different segments of PE
backbones [38]. Copyright 2018 American Society of Mechanical Engineers.
2.3. Inter-chain coupling

with the number-density increase of the side-chains is further shown in Weak inter-chain coupling has been considered as a bottleneck for
Fig. 12(a). A larger number of side-chains could lead to a lower thermal thermal conduction in polymers. Enhancing inter-chain coupling
conductivity for both aligned and fork arrangements of side-chains strength could potentially lead to a higher thermal conductivity of
[38]. With the increase of the number density of side-chains, the polymers. For example, the thermal conductivity of polymer salts (Poly
thermal conductivity of a PE-ethyl chain finally converges to be only (vinylsulfonic acid Ca salt)) could be as high as 0.67 W·m−1 K-1 due to
about 40% that of the pristine PE chain [38]. the relative strong electrostatic forces between the ions in different
A larger length of side-chains could also lead to a smaller thermal polymer chains, in comparison with the common 0.2–0.3 W·m−1 K-1 for
conductivity. Ma and Tian [39] reported that the thermal conductivity most polymers [43]. In addition to the ionic inter-chain coupling, hy-
decreases with the increase of the side-chain length, which is attributed drogen bond (H-bond) is another type of relatively strong inter-chain
to the increased phonon scattering. The thermal conductivity become coupling, which is 10–100 times stronger than the vdW interaction.

Fig. 12. Effects of number density and length of side-


chains on the thermal conductivity: (a) a single PE
chain with different number density of side-chains
[38]. copyright 2018 American Society of Mechanical
Engineers; (b) a single bottlebrush chain with different
side-chain length. Insets show the schematic of poly-
mers [39]. Copyright 2017 American Institute of Phy-
sics.

7
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Therefore, large concentration of H-bonds could also lead to a higher could be explained by the fact that the backbone is mostly-extended at
thermal conductivity. Cross-linking to form chemical bonds is appar- ϕPAP = 0.3 while the H-bond density is above the percolation threshold
ently one of the most effective way in enhancing the thermal con- to form a continuous thermal network (as illustrated in Fig. 13a).
ductivity. Here we discuss the effect of H-bonds on the thermal con- How the H-bonds connect to the backbone has an important effect
ductivity in Section 2.3.1, and then the effect of covalent cross-links in on the thermal conductivity. Fig. 13(c) and (d) show that the thermal
Section 2.3.2. conductivity of the polymer blends are higher than that of either
component polymer, which can be explained by the strong H-bonds
connected closely to the polymer backbones through the low-mass and
2.3.1. Hydrogen bond short chemical linkers as illustrated in Fig. 13(b). On the contrary, the
Hydrogen bond is a strong electrostatic interaction between the thermal conductivity of the PAP-PVPh polymer blends is lower than
proton (or hydrated proton) and the lone electron pair(s) in O, N and F that of either constituent polymers in Fig. 13(e), which is because the
atoms. The H-bonds are beneficial to increase the thermal conductivity, H-bonds not directly connected with backbones as that illustrated in
because they could enhance the inter-chain coupling to form con- Fig. 13(b).
tinuous thermal networks to provide more heat-transfer pathways. In one of the most recent works, the importance in the formation of
Therefore, a large density of H-bonds is usually desired for synthesizing a continuous thermal network is further confirmed. The enhanced
high thermal conductive polymers [44,45]. Polymer blending is an ef- thermal conductivity of polymer blends is attributed to the H-bond-
fective way to increase the density of H-bonds because one component induced enlargement of chain coils and the continuous microstructures
polymer can supply donating group while the other to supply accepting in polymers [52]. As illustrated in Fig. 14, by tailoring the distribution
groups. Biopolymers in which the H-bond molecular group exist widely and the density of H-bonds among polymer blends with PVA and bio-
are usually applied in polymer blends, such as gelatin which possesses polymers (lignin, gelatin), an optimum concentration ratio of polymer
abundant carboxylic, amide, and amine groups [46–48], lignin which blends to enhance the thermal conductivity is also observed. The
owns abundant hydroxyl and aldehyde groups [49,50]. highest thermal conductivity of PVA/lignin and PVA/gelatin blends are
The H-bond density and the mixture ratio of polymer blends could 0.53 W·m−1 K−1 at 2 wt% lignin loading and 0.57 W·m−1 K−1 at 5 wt%
strongly affect the thermal conductivity. To enhance the thermal con- gelatin loading, respectively. Specifically, the thermal conductivity of
ductivity, polymer blends not only need to be mixed uniformly to en- the PVA polymer blended with 10 wt% lignin and 10 wt% gelatin
sure a homogeneous distribution of H-bonds but also to allow polymers reaches 0.71 W·m−1 K−1, which is much higher than that of the PVA/
to intertwine within the radius of gyration to supply a continuous lignin and the PVA/gelatin polymer blends. This high thermal con-
thermal network. This is systematically studied in Ref. [51], by mixing ductivity is attributed to the H-bond networks formed by these three
(poly(N-acryloyl piperidine) PAP with the H-bond donating polymer polymer blends (PVA, lignin and gelatin) to supply more heat transfer
(poly(acrylic acid) (PAA), poly(vinyl alcohol) (PVA), or poly(4-vinyl pathways, as illustrated in Fig. 14.
phenol) (PVPh)). As shown in Fig. 13, there is an optimum composition In addition to bridging different polymer chains to form heat-
of PAP at ϕPAP = 0.3 to enhance the thermal conductivity of polymer transfer networks (bridging effect), H-bonds could also serve as “soft
blends in Fig. 13(c) and (d), where ϕPAP is the mixture ratio of PAP. This

Fig. 13. Thermal conductivity of polymer blends by engineering H-bond interactions: (a) Illustrations of heterogeneous (left) and homogeneous (right) distributions
of H-bonds; (b) Inter-chain H-bonds (dashed lines); (c)-(e), measured thermal conductivities of PAP:PAA, PAP:PVA and PAP:PVPh polymer blends as a function of the
mixture ratio of PAP (ϕPAP ) [51]. Copyright 2015 Nature Publishing Group.

8
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 14. H-bonds in polymer blends and phonon transport for small, large, and continuous coil structures [52]. Copyright 2017 American Chemical Society.

grips” to suppress the segmental rotation of polymers, which reduces example, Tonpheng et al. [64] experimentally achieved a 50% en-
phonon scattering and thus leading to a higher thermal conductivity hancement of the thermal conductivity for the PE polymer at a higher
[53,54]. Through MD simulations of amorphous polymer blends, Wei cross-linking densities. Using MD simulations, Kikugawa et al. [44]
et al. [55] proposed that locally ordered lamellar structures could form reported that the crosslink could significantly increase the thermal
in the polymer blends due to the large inter-chain interactions and thus conductivity of the PE polymer and the increase of the thermal con-
enhancing the thermal conductivity. ductivity is proportional to the crosslink density.
It is not always an effective way to tailor the thermal conductivity It might be intuitive to think that the thermal conductivity en-
by modulating the density and the distribution of H-bonds. hancement is mainly due to the strong covalent bonds. However, recent
Experimental works show that the thermal conductivities of most MD simulations by Rashidi et al. [65] showed that the enhanced
polymer blends are still too low, 0.12–0.38 W·m−1 K−1, and the effect thermal conductivity with crosslinks cannot be explained completely by
of H-bonds on the thermal conductivity is negligible [56]. Numerous only considering covalent bonds. As shown in Fig. 15, the covalent
polymers with strong hydrogen bonding still possess a quite low
thermal conductivity, for example, 0.25 W·m−1 K−1 for nylon-6,6 [57].
The MD simulations by Zhang et al. [58] confirms that polymers with
stronger inter-chain interactions (H-bonds, Columbic interactions, etc.)
do not necessarily have higher thermal conductivity, because of pos-
sible weaker backbone bonds caused by inter-chain interactions. It was
also revealed that H-bonds may increase the inter-chain phonon scat-
terings and thus suppressing the contributions of acoustic phonon
modes to the thermal conductivity [59]. Similar effect of vdW inter-
actions on the thermal conductivity was also observed in Ref. [60],
where the thermal conductivity of PE strands decreases weakly as the
chain number increases, because vdW interactions between chains in-
troduce slightly more phonon scattering. Huang et al. [61] even at-
tributed an exceptionally high thermal conductivity (up to
416 W·m−1 K−1, comparable to copper) of the dragline silk of spider to
the break of the inter-chain H-bonds, which is needed for restructuring
polymer chains.

2.3.2. Crosslink
Crosslinks can form efficient heat conduction pathways and net- Fig. 15. Thermal conductance in two PMMA chains cross-linked by CH2, nor-
works by connecting polymer chains with strong covalent bonds. It is malized by that with a cross-link density of 0, where the thermal conductance is
natural to expect that the thermal conductivity would increase with the defined as the heat flow divided by temperature difference [65]. Copyright
increasing number of crosslinks in the polymer network [62,63]. For 2017 American Chemical Society.

9
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

bonds contribute to only 20% of the total thermal conductance for all However, the known high thermal conductivity of individual CNT
number density of cross-links. More interestingly, the nonbonding in- and GNP does not easily translate to high thermal conductivity nano-
teractions not only contribute to the majority of total conductance, but composites. Until now, the thermal conductivity of CNT-polymer na-
also their contribution of the total conductance increases with the in- nocomposite is still much lower than that with the commonly used
creasing density of crosslinks. In addition to the covalent bonds con- metallic or ceramic fillers like aluminum or silicon carbide. There could
necting different chains, another effect of the crosslinks is to bring the be two possible reasons: 1). The quality of individual CNT and GNP
polymer chains closer to each other. As a result, the non-bonding used in nanocomposites can be quite different from those prepared for
coupling (vdw, Coulombic or H-bonds) becomes stronger when there individual characterization of thermal properties. CNTs usually possess
are more crosslinks in the polymer, which in turn significantly enhances some defects which could greatly reduce its effective thermal con-
the thermal conductivity. ductivity [115], such as Stone-Wales defect [116], doping or vacancy
Interestingly, crosslinks could also counter-intuitively decrease the defects [117], and inter-tube junction [118,119]. Besides the intrinsic
thermal conductivity of polymers. For example, Yu et al. [66] experi- defects of CNT, the kinks, twists and waviness formed in the CNTs could
mentally showed a 30% reduction in the thermal conductivity of HDPE, also reduce the aspect ratio of CNT, and thus lead to a lower thermal
because the cross-links in HDPE suppressed the crystallization of conductivity of CNT-in-epoxy composites than expected [120,121].
polymer molecules. Although the crosslinks could enhance the inter- Similarly, the reduced thermal conductivity due to the waviness is also
chain coupling strength, these crosslinks could also break the periodi- observed in GNP [122]. To remove such defects in CNTs and GNPs,
city along the polymer chains, which could result in strong phonon surface modification methods have been developed to alter the mor-
scattering. For example, MD simulations performed by Ni et al. [67] phology and defect density, such as acid [123] and plasma treatment
suggest that a 10% crosslinking in PE polymer could result in a 44.2% [124–126]. 2). Interface thermal resistance across the nano-filler and
reduction in the thermal conductivity along the chains. How to en- polymer can greatly reduce the benefits from the high thermal con-
gineer the cross-link to tune the thermal conductivity of a polymer is ductivity nano-fillers. When the filler concentration is low and a filler
still an open question. network is not formed, the large interfacial thermal resistance between
fillers and matrix usually results in a low thermal conductivity. Fur-
3. Thermal conductivity of polymer nanocomposites thermore, the polymer molecules wrapping around the CNT or GNP
could induce strain and shape change and thus reducing the thermal
In addition to engineering polymer chains and morphology at the conductivity of the CNT and GNP [127].
atomic/molecular level, the thermal conductivity of polymers can also Increasing the amount of fillers to form a network can significantly
be enhanced by adding highly thermal conductive fillers. Although the enhance the thermal conductivity [128–132,134,135]. However, a high
composites incorporated with macro- or micro-fillers have already been concentration of fillers could compromise other important properties of
widely studied for close to a century using the effective medium theory a polymer, such as mechanical, electrical, and optical properties. Con-
[68–71], the thermal conductivity of composites filled with nano-fillers structing a 3-dimensional (3D) filler network by uniformly distributing
can be quite different and not well understood yet. Different from the fillers is a promising method to improve the thermal conductivity with
macro- or micro-scale composites, the large specific surface area of a relatively low concentration of fillers. However, there yet exists a
nano-fillers can lead to large contribution of interfacial thermal re- clear theoretical understanding of heat transfer mechanism in nano-
sistance in a nanocomposite. In addition, it is very challenging to composites with 3D filler networks. Most of the theoretical models are
mathematically describe the heat conduction if nano-fillers form a based on the effective medium theory (EMT) [136–142] or the perco-
network in a nanocomposite. The thermal conductivity is determined lation-based theory [143–145], which might not be useful for nano-
not only by the polymer matrix and the fillers, but also the interaction composites with 3D filler network. Many of the numerical methods,
between filler and matrix and among fillers (the filler network). In the such as Monte Carlo method [146–151], finite element method
following, we firstly introduce the influence of fillers on the thermal [152,153] and lattice Boltzmann method [154,155], recently devel-
conductivity, and then discuss the thermal transport mechanism of oped by the nanoscale heat transfer community might shed some lights
different filler networks. in the physical understanding, while challenging to be used for the
The commonly used fillers in nanocomposites can be categorized design. In the following, we discuss the polymer nanocomposites both
into the metallic, ceramic and carbonous fillers. The metallic fillers, without inter-filler network and with inter-filler network to highlight
such as copper nanoparticles [72,73], copper nanowires [74,75], alu- the current understanding on the thermal conductivity of polymer na-
minum fibers [76], silver particles [77], gold and palladium powders nocomposites.
[78,79] could be used to enhanced the thermal conductivity of a
polymer nanocomposite. However, these metallic fillers may also lead 3.1. Nanocomposites without an inter-filler network
to an increase of electrical conductivity, which prevents their applica-
tions with electrical insulation requirement. Although the electrical When there is no inter-filler network formed in a polymer nano-
conductivity of metallic fillers could be somewhat tailored by oxidation composite, most of the thermal resistance comes from the polymer
or surface treatment [80,81], the high thermally conductive ceramic matrix and the interfacial thermal resistance between matrix and fillers.
fillers is more preferable for not only their electrical insulation property In this case, the thermal conductivity increases gradually with the in-
but also thermal stabilities. Typical high thermal conductivity ceramic creasing concentration of fillers, and usually does not exhibit a perco-
nano-fillers are magnesium oxide (MgO) [82–84], aluminum oxide lation behavior [156,157]. The concentration-dependent thermal con-
(Al2O3) [85–88], silicon nitride (Si3N4) [89–91], silicon carbide (SiC) ductivity of a nanocomposite without inter-filler networks could be
[92–94], zinc oxide (ZnO) [95], aluminum nitride (AlN) [96–99], and separated into three regimes. In the regime of low filler concentration,
boron nitride (BN) [100–104]. Compared with the metallic and ceramic the fillers are independent from each other, thus the thermal con-
fillers, nanostructured carbon fillers have attracted more intensive in- ductivity increases with the increasing concentration of fillers (first
terests because of their high thermal conductivity. For example, ex- rise). However, with the further increase of filler concentration, the
panded graphite (EG) has a thermal conductivity of about fillers tend to aggregate because of insufficient volume for a high
300 W·m−1 K−1 [105], graphene nanoplate (GNP) possesses a thermal concentration [158]. When the fillers aggregate, the contact area be-
conductivity as high as 1000–5000 W·m−1 K−1 [106–110], and CNTs tween the filler and the polymer matrix decreases dramatically, and the
are regarded as the most promising candidates owing to their high thermal conductivity also sharply decreases. Besides the interfacial
mechanical strength, chemical stability and high thermal conductivity thermal resistance, the changing morphologies of conducting fillers (i.e.
of 1000–3000 W·m−1 K−1 [111–114]. long aspect ratio fillers such as nanowires and CNT can be bended) also

10
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

exert negative effect on the thermal conductivity of fillers, thus redu- matrix, a high filler loading concentration will suppress the crystal-
cing the effective thermal conductivity of the nanocomposites lization of matrix and thus leading to a decrease of the thermal con-
[170,159]. If the filler concentration is further increased, a second rise ductivity of nanocomposites [180,181]. Recently, the research of Ding
of the thermal conductivity could occur because the clusters formed by et al. [181] showed that the thermal conductivity of GNR@PA6 nano-
aggregation of fillers could contact with others to form heat transport composites slightly decreases with the increasing concentration of GNR
pathway. fillers when GNR concentration was higher than 0.5 wt%. By using X-
The thermal conductivity in the ‘first rise’ regime has been well ray diffraction and differential scanning calorimetry to analyze the
studied and reasonably well understood, with the EMT, under the dilute morphology of GNR@PA6, Ding et al. unveiled that a low concentration
limit, by taking into account of interfacial thermal resistance of GNR plays a role in the heterogeneous nucleation to improve the
[68,160,161]. At a low filling concentration, the heat flux lines in a crystallization rate of PA6 and thus enhancing the crystallinity degree,
nanocomposite generated by one particle are not distorted by the pre- while high concentration of GNR would obstacle the crystallization.
sence of the neighboring particles because the distance between The size of fillers could greatly affect the thermal conductivity of a
neighboring particles is much larger than their size. However, for polymer nanocomposite through influencing the contact area between
higher particle concentrations the distance between neighboring par- the fillers and the matrix. Noh et al. [182] showed that larger fillers
ticles can be of the order of the particle size or smaller and the inter- lead to an obvious thermal conductivity enhancement due to the in-
action among particles have to be considered, which results in distor- creased interface conductance. Kim et al. [183] reported a similar result
tion in heat flux that is different from the prediction of the single that a larger lateral size and thickness of the GNP may decrease the
particle assumption. To take into account the particle interactions, contact resistance between GNP and polymer matrix to improve the
based on the differential effective medium theory [162], Ordonez- thermal conductivity of nanocomposites. Yu et al. [184] even showed
Miranda et al. [163–165] have developed a crowding factor model to that the filler smaller than a critical size cannot enhance the thermal
predict the thermal conductivity of composites, where the crowding conductivity because the large interfacial thermal resistance counteract
factor is determined by the effective volume fraction of fillers. Their the contribution of the high thermal conductivity of fillers. However,
model generalize other EMT models which are suitable for nano- one needs to be careful when choosing the filler size, because other
composites, but also give a good thermal conductivity prediction of physical properties could be strongly affected when the fillers are too
nanocomposites with a high filling ratio of fillers. large.
As an example, the experimental thermal conductivities of The size of fillers could also greatly affect the thermal conductivity
MWCNT@PP are shown in Fig. 16 [166]. The thermal conductivity through shortening the mean free path of energy carriers in a nano-
exhibits a non-monotonic trend as a function of the filler concentration. composite incorporated with randomly-distributed fillers. The EMT has
The first rise of thermal conductivity is attributed to a uniform filler been modified to take this effect into account by considering the energy
distribution, while the aggregation dominates the following decrease of carrier collision cross-section of the fillers and the average distance that
the thermal conductivity with the increase of volumetric loading of the energy carriers can travel inside the fillers [185,186]. The modified
MWCNT. With 3 vol% MWCNT loading in the matrix, the thermal EMT could give good predictions compared to the numerical ap-
conductivity decreases because some MWCNTs have to bend or ag- proaches based on the BTE, Monte Carlo simulations and experiments.
glomerate to fit into the small packing volume. Similar decrease trend Further development of modified EMT models has taken into account
of thermal conductivity is also observed in Refs. [167–170], which is the spectral phonon properties including MFP, polarizations and wave
attributed to the increased interfacial thermal resistance caused by vectors [187–189].
aggregation. The second rise of the thermal conductivity may result In short, when no inter-filler networks are formed, the aggregation
from the heat transport pathway formed among clusters which are of fillers and the large interfacial thermal resistance play important
formed by aggregation of fillers. roles in the thermal conductivity of polymer nanocomposites
From the first rise to the decrease in the effective thermal con- [131–135]. Here we introduce the methods to reduce the interfacial
ductivity, there exists a critical filler-matrix interfacial thermal re- thermal resistance and to prevent the aggregation of nano-fillers.
sistance determining the turning point of the thermal conductivity. For
example, the computational study of SWCNT@polymer shows that this 3.1.1. Methods of enhancing interface conductance
critical SWCNT-matrix interfacial thermal resistance (Rc) is about The surface treatment and functionalization of nano-fillers has been
2 × 10−7 m2KW−1 [171], as illustrated in Fig. 17. If the interfacial widely implemented to enhance the filler-polymer matrix coupling to
thermal resistance is smaller than Rc, the effective thermal conductivity reduce the interfacial thermal resistance [191–194]. We summarize the
of the nanocomposite would always increase with the volumetric ratio major practices according to the type (dimensionality) of fillers.
due to the large thermal conductivity of the SWCNTs. When the inter-
facial thermal resistance is larger than Rc, the thermal conductivity
decreases with the increasing concentration of fillers. Unfortunately,
the existing acoustic mismatch model and the diffuse mismatch model
[172] could not accurately estimate the interfacial thermal resistance
between the filler and the polymer matrix. MD simulations have been
applied to elucidate the interfacial thermal resistance [173,174], where
only the bonding or adhesion strength across the interface is con-
sidered. In a real nanocomposites, the interface thermal resistance be-
tween nano-filler and polymer can be dependent on many other factors
including voids and molecule line-up at the interfaces.
The addition of nano-filler can also affect the crystallization of the
polymer matrix. Both SWCNT and MWCNTs can act as nucleation
agents in increasing the polymer crystallization rate, which results in a
larger thermal conductivity of polymer matrix and a reduced interfacial
thermal resistance [175–177]. The crystallization behavior and the
thermal conductivity of polymer matrix could be also enhanced by
other kind of fillers, such as nanoparticles [178,179]. While a low Fig. 16. Experimental thermal conductivities of MWCNT@PP composites as a
concentration of fillers could improve the crystallization of polymer function of MWCNT volume fraction [166]. Copyright 2016 Elsevier Ltd.

11
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

could also increase the thermal resistance of the matrix. As shown in


Fig. 19, the functionalization increases the thermal resistance of the
near-interface polymer layer (R2) , until R2 saturates when the functio-
nalization degree (x ) becomes larger than 0.1. The functionalization
significantly reduce the interfacial thermal resistance R1 until x reaches
0.15, while a further increase of x will not reduce the interfacial
thermal resistance. The result of R1 + R2 reaches a lowest value at about
x = 0.05, which suggests that increasing the functionalization degree
cannot always lower the total effective thermal resistance [206].
2) 2D fillers (graphene and boron nitride)
Similar to 1D fillers, the interfacial thermal resistance between 2D
fillers and polymer matrix is also the bottleneck for achieving high
thermal conductivity in the nanocomposites. For example, the interface
conductance between graphene and poly(methyl methacrylate)
(PMMA) is as high as that between CNT and polymers, which is about
1.906 × 10−8 m2KW−1 [207]. To reduce the large interfacial thermal
resistance between the 2D-fillers and the matrix, the surface functio-
Fig. 17. Thermal conductivity enhancement as a function of SWCNT-matrix nalization of 2D fillers, such as graphene, and BN nanosheets has been
interfacial thermal resistance (Rc) for SWCNT/polymer composites at SWCNT widely studied. A strong covalent bonding of BN nanosheets on epoxy
loadings of 1 vol%, 3 vol%, 9 vol% and 22 vol%. km denotes the bulk thermal matrix could result in a higher thermal conductivity [208].
conductivity of the polymer matrix [171]. Copyright 2012 Elsevier Ltd. However, surface functionalization of 2D fillers does not always
lead to a smaller interfacial thermal resistance. The effect of surface-
1) 1D filler (CNT) grafted chains on the thermal conductivity of graphene@polyamide-6.6
Although CNT possesses a very high thermal conductivity, the nanocomposites has been studied using MD simulations [209]. It turns
thermal conductivity of CNT@polymer composites is generally lower out that the thermal conductivity perpendicular to the graphene plane
than 10 W·m−1 K−1. The challenge of achieving high effective thermal is proportional to the grafting density, while the in-plane thermal
conductivity primarily originates from the large interfacial thermal conductivity of graphene drops sharply as the grafting density in-
resistance between the CNT and the surrounding polymer matrix [195]. creases. There exist an intermediate grafting density for maximal en-
The interfacial thermal resistance between CNT and polymer matrix can hancement in the thermal conductivity of nanocomposites, as shown in
range from 0.1 × 10−8 to 15 × 10−8 m2·K·W-1 [196–199]. When heat Fig. 20. Besides the optimal grafting density, there also exists an op-
flows in a CNT composite, the majority of the temperature drop occurs timal balance between grafting density and grafting molecular length to
at the interface while the temperature in CNTs essentially remains obtain the maximum enhancement.
uniform [199,200]. In addition to surface functionalization using organic molecules,
Many efforts have already been made to reduce the interfacial inorganic MgO has also been successfully coated on the surface of
thermal resistance between CNTs and polymer matrix through func- graphene using a simple chemical precipitation method to enhance the
tionalization. MD simulations by Huang et al. [201] showed that the interfacial thermal conductivity, where MgO served as an effective in-
interfacial thermal conductance of functionalized CNTs is enhanced by terface to strengthen the interfacial adhesion between the MgO and the
100% compared with the pristine CNT-polymer interface, because of epoxy matrix [210]. Non-covalent modification of h-BN nanoparticles
the enhanced coupling between CNT and the matrix. Kaur et al. [202] with PDA could bring about a higher thermal conductivity of bisphenol-
also reported a six-fold reduction of interfacial thermal resistance be- E-cyanate-ester based composites than that without surface functiona-
tween matrix and MWCNT arrays by bridging the interface with short, lization [211]. By depositing silver nanoparticles on the BN nanosheets,
covalently bonded organic molecules. A COOH-functionalization was Wang et al. [212] showed that the epoxy composite possess a higher
found to reduce the interfacial thermal resistance, resulting from the thermal conductivity than that with pristine BN nanosheets due to the
modified CNT-matrix interface [203]. Besides functional groups cova- bridging connections of silver nanoparticles among BN nanosheets.
lently bonded on CNTs, the thermal conductivity could also be en-
hanced by incorporating silver nanoparticles onto the CNT surface, due
to the superior thermal conductivity of silver and the reduced Ag-CNT
interfacial resistance [204].
Although covalent functionalization of nano-fillers could reduce the
interfacial thermal resistance, they could also reduce the thermal con-
ductivity of fillers. MD simulations by Ni et al. [205] showed that
functionalizing aromatic polymer HLK5 (C22H25O3N3) onto the CNT
can efficiently decrease the interfacial thermal resistance between the
CNTs and different types of polymer matrices (PS, epoxy, and PE).
However, the thermal conductivity of CNTs can be also decreased by
functionalizing HLK5, octane, or hydroxyl on CNTs, as shown in Fig. 18.
It shows that a larger functionalized surface of CNTs could lead to a
smaller effective thermal conductivity of the polymer composite.
When one end of a molecule group is functionalized on the surface
of a CNT and the other end on the polymer matrix, this kind of surface
functionalization is also called cross-link. Although the cross-links
could improve the thermal coupling between the CNTs and the polymer
matrix, they could also suppress the thermal conductivity of CNTs. The
thermal conductivity of CNTs can be reduced to half by cross-links
compared with that of the freestanding CNTs [206]. The cross-links Fig. 18. Thermal conductivity of functionalized CNTs scaled by that of pristine
CNT [205]. Copyright 2015 American Chemical Society.

12
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

increasing concentration of MWCNTs when its concentration is larger


than 1 wt%, while the thermal conductivity of BN-c-MWCNT@PI keeps
increasing with the increase of BN-c-MWCNT concentrations. Such
different results originate from the different dispersibility of MWCNTs
and BN-c-MWCNT in the PI matrix. The dispersibility of MWCNTs is
poor, and the aggregation happens when its concentration is higher
than 1 wt%, which increase the interfacial thermal resistance between
MWCNTs and polyimide. For BN-c-MWCNT@PI composites, the fillers
do not aggregate because of the BN coating. Similar phenomenon was
also observed in Ref [221]. that by coating PVP on the surface of silver
nanowire fillers could hinder their aggregation in a silver-nanowire@
epoxy nanocomposite.
Unfortunately, it remains very challenging to simultaneously
achieve strong filler-matrix coupling and low aggregation. Surface
functionalization for suppressing filler aggregations usually leads to a
weak filler-matrix coupling, making it difficult to achieve high thermal
conductivity [222]. In addition to the surface treatment and functio-
Fig. 19. Interfacial thermal resistance between CNTs and the PE polymer (R1), nalization, dispersant is another choice to prevent aggregation of fillers.
and the thermal resistance of the near-interface polymer layer (R2). The func- graphene oxide (GO) appears to be a good dispersant, as it has similar
tionalization degree is defined as the ratio of the number of functional groups to lattice structure to graphene and could also lead to a stable dispersion
the number of carbon atoms in the outermost layer of the CNT. Inset shows
of CNTs in aqueous media through the π-π stacking interaction between
topologies of the pristine (5,5) SWCNT nanocomposite and the functionalized
GO and CNTs [223,224]. Clay could also promote a uniform distribu-
one with x = 0.1 [206]. Copyright 2015 Elsevier Ltd.
tion of CNTs to form a percolated network structure in a polymer matrix
[224,225]. Polyvinylpyrrolidone (PVP) is helpful to improve the dis-
persion of CNTs and induce denser CNT network structure in poly(vi-
nylidene fluoride) PVDF matrix [226].
Some special preparation process could also improve filler dis-
tribution. The thermal conductivity of the BN@PE composites has been
improved after multistage stretching extrusion process, because the
aggregation of BNs were reduced due to the strong shear field [227]. On
the contrary, rheological studies showed that the shear could also in-
duce an aggregation of CNTs, and CNTs with a longer aspect ratio
possess a larger shear-induced aggregation [228]. The shear-induced
distribution of fillers depends on not only the filler but also on the
processing details. Some practical and feasible preparation are still
needed to improve the distribution of CNT and other fillers in a polymer
matrix.
The orientation of fillers with large aspect ratio is another crucial
factor determining the effective thermal conductivity of nanocompo-
sites. Similar to the polymers, the alignment of filler orientation can be
improved by tension [229,230], templating [231,232] and
Fig. 20. Thermal conductivity of nanocomposites scaled by that of the polymer
matrix. f is volumetric fraction of graphene fillers [209]. Copyright 2016
American Chemical Society.

3.1.2. Improvement of filler distribution


In addition to the interfacial thermal resistance between the fillers
and the matrix, the thermal conductivity enhancement could also lim-
ited by the tendency of filler aggregation. To suppress the filler ag-
gregation, several methods have been developed to improve filler dis-
tribution in the past few decades including surface functionalization,
addition of dispersant, and special preparation techniques.
CNTs tend to aggregate together because of the strong vdW force
and the chemical inertness caused by their unique sp2 bonding
[213,214]. To prevent aggregation of CNTs, surface modification of
CNTs through covalent or non-covalent approaches have been devel-
oped [215,216]. Although chemical functionalization of the CNTs im-
proves the dispersion of CNTs, the covalent functionalization can also
distort the structure of CNTs and thus sometimes even reducing the
thermal conductivity [217]. Coating CNTs with inorganic materials
such as BN or alumina has also been reported to significantly improve
the dispersibility of CNTs [218–220]. The thermal conductivities of
polyimide (PI) based nanocomposites incorporated with MWCNTs or
Fig. 21. Thermal conductivity of MWCNT@PI and BN-c-MWCNT@PI. Insets
BN-coated MWCNT (BN-c-MWCNTs) are compared, as shown in Fig. 21
are the transmission electron microscope images of MWCNTs and BN-c-
[218]. The thermal conductivity of MWCNT@PI decreases with the
MWCNTs [218]. Copyright 2014 Royal Society of Chemistry.

13
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

electrospinning [233]. For example, Datsyuk et al. [233] prepared a found to be using the prolate ellipsoids compared to other shapes like
high thermally conductive CNT@polybenzimidazole nanocomposite by oblate ellipsoids and spheres, because it is easier to form a conductive
electrospinning. The in-plane thermal conductivity of the composite network with prolate ellipsoids [241–243]. Although 2D fillers are
reaches 18 W·m−1 K−1 at 1.94 wt% CNT concentration, 50 times larger better than 1D fillers for enhancing the thermal conductivity because of
than the polymer matrix. This high thermal conductivity is attributed to the face-contact between 2D fillers, the 1D filler is more apt to form a
the excellent CNT alignment in the polymers. heat-transport inter-filler network at a lower concentration which in
Electromagnetic field or electric field could be applied to align fil- turn might enhance the thermal conductivity due to their large aspect
lers. For example, Lin et al. [234] utilized an external magnetic field to ratio. The percolation threshold for 2D and 0D fillers is much higher
align hexagonal boron nitride (h-BN) to obtain a high thermal con- than that of 1D fillers because of different aspect ratios where the
ductivity along the alignment direction. Cho et al. [235] used a high theoretical percolation threshold for heat transport is proportional to
direct current electric field to enhanced thermal conductivity in the out- the reciprocal of aspect ratio [250]. Due to the high aspect ratio of CNTs
of-plane direction of the BN@polysiloxane nanocomposites. The ice (up to 103–104), it is possible to establish percolation paths at a low
templating self-assembly strategy is a proven effective method to form CNT concentrations, usually lower than 1 vol%, to enhance the thermal
well-aligned fillers along the ice-growth direction [236–238]. The conductivity. It has been shown that enlarging the aspect ratio of CNT
thermal conductivity of BN@epoxy-resin nanocomposites prepared to reduce the number of contacts required to form a percolating net-
with this strategy is as high as 4.42 W·m−1 K−1, much higher than that work could improve the thermal conductivity of the composite [158],
of nanocomposites filled with randomly distributed BN which is less and the thermal conductivity of CNT@polymer composites can be ef-
than 1.81 W·m−1 K−1 [239]. fectively controlled by adjusting the length of the CNT fillers [251]. On
the other hand, the percolation threshold for 2D and 0D fillers are
3.2. Nanocomposites with inter-filler network usually much higher than 1 vol%. For example, the percolation
threshold of the thermal conductivity in PE based nanocomposites in-
It is natural to expect that an inter-connected network can be corporated with graphite powder is as high as 10 vol% [252].
formed when the filler concentration is higher than a critical value, i.e. The research carried out by Zhao et al. [253] is introduced here to
the percolation threshold [240]. Although a single kind of fillers can illustrate the thermal conductivity change with a percolation phe-
form a 3D network, there still remains lots of voids that might not be nomenon when an inter-filler network is formed, as shown in Fig. 22.
filled tightly by the polymer matrix. The introduction of another kind of Three different fillers, raw carbon fiber (raw-CF), copper-coated CF
fillers is beneficial to fill the vacancies and to further enhance the ef- (Cu-CF) and (3-mercaptopropyl)trimethoxysilane-Cu-CF (M–Cu-CF),
fective thermal conductivity of nanocomposites. Incorporating with were incorporated in silicone rubber (SR) matrix to enhance the
only a small amount of hybrid fillers could lead to an equally increased thermal conductivity. These polymer composites all exhibited an en-
thermal conductivity with respect to that filled with single fillers. hanced thermal conductivity with a percolation threshold at the filler
However, the thermal conductivity of nanocomposites even with the concentration of about 1 wt %. The larger thermal conductivity of na-
inter-filler network is still too low compared to many high thermal nocomposites filled with Cu-CF and M–Cu-CF than that with CF is due
conductivity inorganic materials, due to the large inter-filler thermal to the fiber surface treatment. When the filler loading is less than the
contact resistance. Recently, porous 3D fillers, such as carbon foam and percolation thresholds, the thermal conductivities increase slowly, be-
graphene foam, have drawn great interest because of their intrinsic 3D cause of the large interfacial thermal resistance between fillers and
network structure with no thermal contact resistance. In Sections matrix. When the filler loadings increased up to about 4.0 wt%, there is
3.2.1–3.2.3, we discuss three kinds of nanocomposites with 3D net- a dramatic increase of the thermal conductivity resulting from the
works. networks. Nevertheless, the thermal conductivities increase slowly
again when the filler concentration becomes larger than 4.0 wt%, be-
3.2.1. Single fillers cause the thermal conductive paths tend to saturate. Besides the widely
The filler network could be greatly influenced by the dimensions observed percolating behavior of polymer nanocomposites in-
and shapes of individual fillers, and thus the thermal conductivity of corporated with 1D-fillers where the increase of thermal conductivity
polymer nanocomposites [241–244]. In a nanocomposite filled with a follows a critical power law [253–255], it has also been widely ob-
single kind of fillers, the influence of filler dimensionality on the served in many experiments that a small amount of 2D fillers filled in a
thermal conductivity (k ) generally follows the trend of k2D > k1D > k 0D ,
where 2D, 1D and 0D stand for nanocomposites incorporated with 2-
dimentional, 1-dimentioanl and 0-dimensional fillers respectively, be-
cause the total contact area per volume increases with the increase of
filler dimensionality [245–247]. Furthermore, the contact area can
have more influence on the thermal conductivity of fillers with lower
dimensionality than larger dimensionality. The defects such as kinks
and twists formed in each individual CNT could lead to a reduction of
its thermal conductivity, while defects in 2D fillers have no such im-
portant influence. For example, platelet-like 2D fillers are preferred for
enhancing the thermal conductivity than a 1D nanowires and nano-
tubes, because of a lower thermal contact resistance of a surface than a
point contact resistance. Among 2D GNP, 1D CNT, and 0D super-full-
erene, the GNP-enhanced PVDF composites possess the highest thermal
conductivity (2.06 W·m−1 K−1), which is about 10-folds enhancement
compared to that of the pristine PVDF [248]. The GNP@epoxy shows a
thermal conductivity of 0.47 W·m-1 K-1 with an increase about 126.4%
compared to that of the neat epoxy, while MWCNT@epoxy shows a
thermal conductivity of 0.33 W·m-1 K-1, an increase for only about 60%
[249]. Fig. 22. Experimental thermal conductivity of raw-CF@SR, Cu-CF@SR, and
The shape of fillers also plays an important role. Among different M–Cu-CF@SR nanocomposites [253]. Copyright 2016 Springer International
shapes of 2D fillers, the highest thermal conductivity enhancement is Publishing AG.

14
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

polymer matrix could result in a significant enhancement of the thermal 3.2.2. Hybrid fillers
conductivity [256,257]. Although a single kind of fillers can form a 3D network, there still
Among 0D, 1D and 2D fillers, the nanocomposites incorporated with remains lots of voids that can hardly be filled by the polymer matrix.
1D fillers are most well-studied. The heat conduction through the net- The introduction of another kind of fillers is beneficial to fill the voids
work in a nanocomposite filled with 1D fillers is influenced by many and to further enhance the effective thermal conductivity of nano-
factors such as inter-filler thermal contact resistance, intrinsic thermal composites. There is always a preferable concentration ratio between
resistance of the fillers, volumetric fraction, aspect ratio of fillers, and different constituent fillers to enhance the thermal conductivity
orientation distribution of fillers. Most recently, a general thermal [259–261]. By studying the thermal conductivity of polycarbonate (PC)
conductivity model for nanocomposites with 1D-filler network has been based nanocomposites filled with MWCNTs and EG, Zhang et al. [262]
developed to include the effects of inter-filler thermal contact re- showed that the maximum thermal conductivity could be achieved with
sistance, intrinsic thermal resistance of fillers, volume fraction of fillers, loading weight ratio of 9:1 between the EG/MWCNTs, and this max-
and orientation distribution of fillers [258]. To describe the competing imum thermal conductivity is higher than that of the nanocomposite
effect of the inter-filler thermal contact resistance and the filler intrinsic with either single kind of fillers. However, there yet exists a theory to
thermal resistance, Zhao et al. [258] defined a dimensionless Biot find the optimal concentration ratio. More studies are still needed to
number to describe this competing effect, probe the relationship between the best concentration ratio and the
geometric properties of hybrid fillers like sizes and dimensionality. It
〈Nc 〉 hL should be also noted that the optimal concentration ratio of hybrid
BiT =
k 0 A0 (1) fillers not always ensure a maximum thermal conductivity, whereas a
much higher concentration of fillers may destroy the synergistic effect
where Nc is the average number of contact junctions for a single fiber of hybrid fillers [263].
with its neighboring fibers, h is the thermal conductance of the inter- According to different structures of filler networks, the polymer
filler contact, k 0, L and A0 is respectively the thermal conductivity, nanocomposites incorporated with hybrid fillers can be categorized into
length and cross-sectional area of the 1D fillers, as shown in Fig. 23(a). four cases: (1) both of two fillers uniformly disperse in the nano-
The thermal conductivity of polymer nanocomposites with filler net- composite without a network because of a low filler concentration
works could then be derived as [258], (without network); (2) one of the fillers forms a network while another
filler remains uniformly dispersed (single network, no synergistic ef-
k BiT
= ns 〈cosθ〉 fect); (3) while one of hybrid fillers distributes uniformly, another filler
k0 2 cosθ L/ H + BiT (2) bridge them together to form a network (single network with sy-
nergistic effect); (4) both kinds of fillers form network, and double
where ns is number of fillers across an arbitrary cross-section with unit
networks are cross-linked (double network with synergistic effect).
area, 〈H 〉 is the average center-to-center distance of two connected
Here we focus on the synergistic effects, i.e., cases 3 and 4, which might
fillers in the heat transfer direction, θ is the angle between the axial
be inspiring to develop new methods to realize the synergetic en-
direction of the filler and the direction of heat transfer.
hancement of thermal conductivity.
It can be shown from Eq. (2) that a larger θ could lead to a higher
1) Single network with synergistic effect
thermal conductivity of polymer nanocomposites. Similarly, the filler-
The single filler network with different hybrid fillers are schemati-
orientation-improvement methods discussed in Section 3.1.2 may be
cally shown in Fig. 24(a)–(c), which are formed with 0D + 1D hybrid
also useful for improving the thermal conductivity of nanocomposites
fillers, 0D + 2D hybrid fillers and 1D + 2D hybrid fillers, respectively.
with filler networks, although such kind of studies has yet been carried
(a) With 0D + 1D hybrid fillers
out. Reducing the inter-filler thermal contact resistance is helpful in
In the nanocomposite incorporated with 0D + 1D hybrid fillers, a
increasing the effective thermal conductivity. The increasing trend of
single network with synergistic effect could be formed with one of
thermal conductivity with increasing volume fraction of fillers greatly
hybrid fillers distributed uniformly and another bridging them together.
depends on the BiT , as shown in Fig. 23(b). For the nanocomposite filled
The 0D + 1D hybrid fillers contribute considerably to the formation of
with CNTs, there is a large inter-filler thermal contact resistance
a more efficient percolating network for the thermal conduction and
(BiT ≪ 1) and the thermal conductivity is sensitive to the volumetric
thus improving the thermal conductivity, compared to that of single
fraction of fillers.
fillers. For example, the thermal conductivity of PVDF + PS polymer

Fig. 23. Thermal transport in a nanocomposite filled with 1D fillers: (a) Schematic of a nanofiber network; (b) thermal conductivity versus volume fraction (Vfe).
Lines showing model results, while circular, square and triangle dots stand for experimental results [258]. Copyright 2017 American Institute of Physics.

15
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 24. Schematic diagrams of single network formed with different hybrid fillers: (a) 0D + 1D fillers; (b) 0D + 2D fillers; (c) 1D + 2D fillers.

blend based nanocomposites incorporated with MWCNTs + SiC hybrid


fillers could be 1.85 W·m−1 K−1 [264], as that shown in Fig. 25.
However, the composite incorporated with only one kind of fillers ex-
hibits a thermal conductivity of only 0.40 W·m−1 K−1 for 2.9 vol%
MWCNTs and 0.98 W·m−1 K−1 for 11.4 vol% SiC. There is an obvious
synergistic effect between MWCNTs and SiC on the thermal con-
ductivity. In this single network with synergistic effect, the MWCNTs
act as heat conducting bridges among the SiC nanoparticles, while SiC
could separate MWCNTs to prevent the aggregation, as that shown in
the inset of Fig. 25.
(b) With 0D + 2D hybrid fillers
The thermal conductivity of poly(3-hydroxylbutyrate) (PHB) based
nanocomposites with a total filler concentration of 50 wt % but dif-
ferent concentration ratio between BN and Al2O3 was studied by Li
et al. [266]. Fig. 26 shows that the maximum thermal conductivity of
BN + Al2O3@PHB is 31% higher than that of the BN@PHB composites
and 196% higher than that of the Al2O3@PHB composites, respectively.
Fig. 25. Thermal conductivity of PVDF + PS blends filled with different types
This suggests a synergistic effect of hybrid fillers on the thermal con-
of fillers. The concentration of MWCNTs is maintained as 2.9 vol % for all
composites. The data of the SiC@PVDF + PS are obtained from Ref. [265]. The ductivity in BN + Al2O3@PHB. Fig. 26(b) shows that with a small vo-
inset schematically illustrates the distribution of MWCNTs and SiC in lumetric fraction of Al2O3 particles, some BN nano-sheets tend to align
MWCNT + SiC@PVDF + PS nanocomposite, where the white part represents along the Al2O3 surfaces to form a network with a greatly increased
the PVDF, the blue part PS, the black lines MWCNTs, and the green circles SiC contact area. The thermal conductivity is therefore improved with the
[264]. Copyright 2013 Elsevier Ltd. concentration increase of Al2O3 particles. When the concentration of
Al2O3 becomes larger than a certain value, the Al2O3 particles form a
network rather than BN which would disperse in gaps among the Al2O3
particles, thus a decrease of thermal conductivity occurs. The optimal

Fig. 26. Experimental thermal conductivity of PHB composites: (a) synergistic effect; (b) effect of hybrid filler concentration ratio, with a total filler concentration of
50 wt% [266]. Copyright 2017 Elsevier Ltd.

16
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

ratio of BN and Al2O3 for obtaining a maximum thermal conductivity is dispersion of CNTs and the formation of denser CNT + GO networks in
43:7 (43 wt% BN and 7 wt% Al2O3), where the maximum thermal the PVDF matrix. Appling same hybrid fillers, a similar synergetic effect
conductivity is about 1.79 W·m−1 K−1. was also realized for PP polymer based nanocomposites where the
(c) With 1D + 2D hybrid fillers highest thermal conductivity was reported to be ∼ 0.35 W·m−1 K−1
Similar synergistic enhancement effect could be achieved with [274].
1D + 2D hybrid fillers. In such kinds of nanocomposites, the CNT is 2) Double networks with synergistic effect
usually applied as the 1D filler, and the BN or GNP is commonly used as Hybrid fillers can form interconnecting double networks, especially
2D fillers. The high rigidity of GNPs can impede their bending in a high with large aspect ratio 1-D and 2-D nano-fillers. The effect of the double
volume fraction and thus preserves their high aspect ratio for providing networks with hybrid 1D + 2D fillers on the thermal conductivity can
more islands to be connected by CNTs. A rigid 2D filler in the hybrid be summarized as [275]: (1) extremely low thermal contact resistance
fillers could also stop bending and coiling of CNTs in the nanocompo- achieved by overlaping interconnections within 2D fillers ; and (2)
site, and thus maintaining the aspect ratio and high thermal con- synergistic effect between 1D-filler network and 2D-filler network
ductivity of CNTs. based on the bridging effect as well as increasing the network density.
He et al. [267] studied the enhancement of polymer-bonded ex- For polymer nanocomposites incorporated with CNT and EG hybrid
plosives (PBX) using the hybrid fillers of 1D CNTs and 2D GNPs. The fillers, double networks are usually formed at a high concentration of
thermal conductivity of PBX based nanocomposites with a filling ratio fillers, as schematically shown in Fig. 28(a). Two interpenetrating
of 1.31 vol% comprising of 10 vol% GNPs and 90 vol% CNTs is about networks are formed instead of just an EG network wrapping CNTs.
1.4 W·m−1 K−1, which is more than twice that of PBX with pure GNPs Such double networks provide more efficient heat conducting paths,
or CNTs. This high thermal conductivity is due to the synergistic effect and a sharp increase of thermal conductivity is expected at just above
of hybrid fillers when GNPs and CNTs could lead to the formation of a the percolation concentration even with only a small amount of CNTs
more efficient 3D percolating network. The synergetic effect of cyanate for HDPE/15 EG/xCNT and HDPE/20 EG/xCNT nanocomposites [276],
ester (CE) resin based nanocomposites is also studied, which is in- as shown in Fig. 28(b). The thermal conductivity of HDPE/10 EG/
corporated with hybrid fillers, including dodecylamine-modified GNPs xCNTs increases linearly with the concentration increase of CNT fillers
(da-GNPs) and γ-aminopropyl-triethoxysilane-treated MWCNTs (f- along the same line by adding the same amount of EG, which suggests
MWCNTs) [268], as shown in Fig. 27(a). The composite with 3 wt% that no synergistic effect exists in this system but only a mix role is
hybrid filler exhibits a 185% increase in thermal conductivity compared obeyed. Wu et al. [277] also reported a sharp increase of the thermal
with that of CE resin matrix, while composites with individual da-GNPs conductivity in PP polymer based nanocomposites via the formation of
and f-MWCNTs exhibits an increase of 158 and 108%, respectively. The double percolated filler network with small-sized MWCNT network
concentration ratio of hybrid fillers also exerts a large effect on the located within loose large-sized EG network. Their results suggested
thermal conductivity, when 3:1 is the optimal concentration ratio of da- that the formation of double networks could effectively reduce the
GNPs and f-MWCNTs for maximizing the thermal conductivity. inter-filler thermal contact resistance and thus significantly increasing
The thermal conductivity of composites filled with hybrid fillers of the effective thermal conductivity of nanocomposites.
BN and MWCNTs is widely studied, and a synergistic improvement in A remarkable synergetic effect with cross-linked double networks
the thermal conductivity was also observed [269–272]. Xiao et al. on the thermal conductivity was also observed in the epoxy-resin based
[270] reported that with a small concentration of CNTs incorporated in nanocomposites incorporating with GNP and MWCNTs hybrid fillers
BN@PVDF composites via melt blending method, the thermal con- [278], as shown in Fig. 29(a). The thermal conductivity of nano-
ductivity of the composite is much higher than that of BN@PVDF composites with 20 vol% CNTs and 20 vol% GNPs could be as high as
composites at the same BN concentration. In addition to the widely 6.3 W·m−1 K−1, which is much higher than that of the composites with
studied hybrid fillers of CNT and BN or GNP, Zhang et al. [273] also 50 vol% CNTs or 50 vol% GNPs alone. The maximum thermal con-
found a greatly improved thermal conductivity in the CNTs + GO@ ductivity is as high as 7.30 W·m−1 K−1, which is about 38 times that of
PVDF nanocomposites with respect to that of CNT@PVDF nano- epoxy-resin matrix. This maximum thermal conductivity is much higher
composites at the same CNT concentration, as shown in Fig. 27(b). It than that when a single network is formed with same hybrid fillers as
was demonstrated that the introduction of GO is beneficial to the shown in Fig. 27, which suggests that the cross-linked double networks

Fig. 27. Experimental thermal conductivity of polymer nanocomposites incorporated with hybrid fillers: (a) cyanate ester resin based nanocomposites filled with da-
GNP/f-MWCNT hybrid fillers [268]. Copyright 2014 John Wiley and Sons; (b) PVDF nanocomposites with CNT + GO hybrid fillers, the filler concentration of GO is
kept at a constant 1 wt%. Inset shows the different dispersion states of CNTs in the CNT@PVDF and the CNT + GO@PVDF composites [273]. Copyright 2015 Elsevier
Ltd.

17
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 28. CNT + EG@HDPE nanocomposite: (a) Schematic illustrating double percolated networks constructed by the hybrid fillers EG and MWCNT; (b) Thermal
conductivity of CNT + EG@HDPE nanocomposites with the concentration increase of CNT while maintaining the EG concentration at 10, 15 and 20 wt% respec-
tively. The 10 EG/xCNT, 15 EG/xCNT and 20 EG/xCNT stand for increasing the concentration of CNT while maintaining the EG concentration at 10, 15 and 20 wt%
respectively [277]. Copyright 2016 Elsevier Ltd.

is much better than a single network to enhance the thermal con- with polymers and the intumescent nature [282–287]. The thermal
ductivity. However, it should be noted that a much higher filling ratio is conductivity of PA6 based nanocomposite with an EG concentration of
usually required to form a cross-linked double network compared with 60 wt% could be as high as 21.05 W·m−1 K−1, approximately 72 times
that of a single network. To reveal the influence of the filling ratio on higher than that of the matrix [286]. Incorporated with ultrathin-gra-
the synergistic effect of hybrid fillers to enhance thermal conductivity, phite foams at a low filling ratio of about 0.8–1.2 vol%, the wax based
the strength of the synergistic effect is defined as (kHYB−kGNP )/ kGNP by nanocomposite could possess a thermal conductivity of about
Huang et al. [278], where kHYB and kGNP is the thermal conductivity of 3.5 W·m−1 K−1, which is 18 times that of the matrix [288].
nanocomposites filled with hybrid fillers and only GNP fillers, respec- To illustrate the dependence of the thermal conductivity on 3D-
tively. The strength of the synergistic effect is shown in Fig. 29(b). It foam structure, microstructures of GF and CF are respectively shown in
shows that the synergistic enhancement of thermal conductivities Fig. 30(a) and (b) [279,281], and the thermal conductivity between GF
happens in the filler concentration ranging from 10 to 50 vol%, and the and CF are compared in Fig. 30(c). It shows that the thermal con-
synergistic effect is more remarkable at a high filler concentration. ductivity depends exponentially on the mass density for both GF and
CF, because there could be more walls for a larger density. The GF with
nanoscale strut-wall is more effective in enhancing thermal con-
3.2.3. 3D foam fillers ductivity than that with microscale strut-walls, because there could be
The self-supported 3D fillers, such as 3D graphene foam (GF) and more heat transfer routes in GF with nanoscale walls at a given GF
carbon foam (CF) can overcome the shortcoming of filler aggregation volume [279]. It is preferable to increase the volume fraction of GF
during the manufacturing process of nanocomposite, and thus pro- through reducing the pore size without increasing the strut wall
viding a more stable 3D thermal transport network to enhance the thickness [279]. By backfilling PMMA into the pores of GFs, 3D thermal
thermal conductivity [279–281]. The worm-like 3D EG is formed when conductive paths could be formed in the nanocomposite at an extremely
the pristine graphite is intercalated with a variety of inserting agents at low concentration of GF, and thus a significant increases of thermal
high temperature. The nanocomposites incorporated with 3D EG fillers conductivity could be achieved (0.35–0.70 W·m−1 K−1) [207]. Li et al.
can be prepared with a melt-blending method due to its good affinity

Fig. 29. Thermal conductivities of epoxy based nanocomposites: (a) synergistic effect of cross-linked double networks with the hybrid filler concentration ratio of 1:1
for CNT + GNP@epoxy nanocomposites; (b) strength of the synergistic effect [278]. Copyright 2012 American Chemical Society.

18
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Fig. 30. Structures and thermal conductivities


of GF and CF: (a) scanning electron microscopy
images of a GF [279]; (b) scanning electron
microscopy images of a CF [281]; (c) thermal
conductivity comparison between GF and CF
[281]. CF1 and CF2 are respectively derived
from Mitsubishi ARA24 and Conoco Dry Me-
sophase. 1000 and 2800 °C is the graphitization
temperature. GF1-6 signifies GF prepared with
different methods. Linear fits to the density
dependence of the thermal conductivity. (c) is
reprint here with the permission of Copyright
2000 from Elsevier Ltd.

[289] also reported that the thermal conductivity of polyamide-6 (PA6) could decrease the thermal conductivity, because they may cause
based nanocomposite filled with 3D GF could be improved by 300% to phonon scatterings and suppress the contribution of acoustic
0.847 W·m−1 K−1 at a GF loading of 2.0 wt%. phonon modes. More studies are still needed to further clarify the
The associated high electrical conductivity of the GF and CF concurrent effects.
sometimes limits their applications in some fields that requires elec- 3) When nano-filler aggregates in the polymer composite, the thermal
trical insulation. The 3D boron-nitride nanosheets (3D-BNNs), which conductivity could exhibit a non-monotonic increasing trend as a
possesses a high thermal conductivity but a quite low electrical con- function of the filler concentration, which is controlled by the ag-
ductivity can be used as an alternative [290–292]. The 3D-BNNs gregation of fillers and the filler-matrix interfacial thermal re-
aerogel exhibits a much better ability to enhance the thermal con- sistance. The thermal conductivity increases at low concentration
ductivity of epoxy resins based nanocomposites, and the thermal con- due to the higher thermal conductivity of nano-fillers. The ag-
ductivity is about 181% higher than that incorporated with BN, due to gregation of fillers reduces the thermal conductivity at an inter-
the unique 3D structure of 3D-BNNs aerogels which has a smaller inter- mediate concentration. If the concentration is further increased, a
BNNs interfacial thermal resistance [290]. By assembling of BNNs on a second rise of the thermal conductivity may occur because the
3D cellulose skeleton, Chen et al. fabricated a cellulose nanofiber-sup- clusters formed by aggregation of fillers could contact with each
ported 3D interconnected BNNs (3D-C-BNNs) [293]. This 3D-C-BNNs other to form heat transport networks.
also has an ultra-high thermal conductivity about 14 times that of 4) To prevent the aggregation of nano-fillers, nano-fillers are often
matrix at a low concentration of BNNs (i.e., 9.6 vol%), which is about functionalized. However, surface functionalization for hindering
3.13 W·m−1 K−1. filler aggregations could also lead to a larger interfacial thermal
resistance. It is highly desirable to simultaneously reduce the in-
4. Summary and outlook terfacial thermal resistance and suppress aggregation of nano-fillers.
5) For nanocomposites filled with uniformly distributed fillers, the
In this review, we focused on the influencing factors and their un- thermal conductivity first increases slowly until the filler con-
derlying physical mechanisms for tailoring thermal conductivity of centration is larger than the percolation threshold, then grows ra-
polymers and polymer nanocomposites. The main research progress pidly and finally converges to a constant value when the filler net-
over the last two decades can be summarized as: work saturates. While it is still a great challenge to form a 3D
network with a small addition of single fillers, hybrid fillers are
1) Improving crystallinity or chain alignment of polymers usually en- usually applied. Incorporating with hybrid fillers, single or double
hances polymer thermal conductivity. Selecting appropriate networks can be formed. Contrasting to the composite with single
polymer species with special chain structures is crucial to further networks, the composite with double filler networks could have a
increase the thermal conductivity. With the influence of the chain much higher thermal conductivity, due to the reduced inter-filler
structure widely studied, it is clear that a larger stretching and thermal conduct resistance and the bridging effect between double
bending strength of backbone bond leads to a higher thermal con- networks. For hybrid fillers, there is always an optimal concentra-
ductivity, while small dihedral-angle strength can significantly re- tion ratio between two fillers for enhancing the thermal con-
duce the thermal conductivity. Besides bond strength of the back- ductivity. Nevertheless, this ratio is still difficult to be determined
bone, a larger weight or number density of side chains give a lower except by experiments, and it is still an open question to completely
thermal conductivity, while the effect of the side chain length may understand how to modulate the effect brought about by hybrid
depend on the morphology of side chains. fillers.
2) The thermal conductivity of pristine polymers could be increased by 6) The 3D fillers could be more effective than hybrid fillers to enhance
enhancing inter-chain coupling, such as through H-bonds and the thermal conductivity because of their intrinsic 3D network
covalent cross-links. The enhancement effect is attributed to the without thermal contact resistance, such as CF, GF, EG and 3D-
following reasons: more inter-chain “thermal bridges”, suppress BNNS aerogel. The thermal conductivity of 3D fillers depends ex-
rotation of polymer chains, and the extended chain coil. On the ponentially on the mass density, because larger density could lead to
contrary, it is also widely reported that the inter-chain coupling more strut-walls.

19
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

Although significant progress has been made over the last two [12] J. Zhao, A.T. Tan, P.F. Green, J. Mater. Chem. C 5 (2017) 10834–10838.
decades in enhancing the thermal conductivity of polymers and [13] H. Eslami, L. Mohammadzadeh, N. Mehdipour, J. Chem. Phys. 136 (10) (2012)
104901.
polymer nanocomposites, there is still a lot of work need to be done, [14] J. Liu, S. Ju, Y. Ding, R.G. Yang, Appl. Phys. Lett. 104 (15) (2014) 153110.
given the technical importance of nanocomposites. Here we list a few [15] C.L. Choy, Polymer 18 (10) (1977) 984–1004.
[16] C.L. Choy, F.C. Chen, W.H. Luk, J. Polym. Sci. B Polym. Phys. 18 (6) (1980)
directions that might be worthwhile for exploration: 1187–1207.
[17] C.L. Choy, Y.W. Wong, G.W. Yang, T. Kanamoto, J. Polym. Sci., Part B: Polym.
1) The enhancement of the intrinsic thermal conductivity of polymers Phys. 37 (23) (1999) 3359–3367.
[18] J. Ma, Q. Zhang, Y. Zhang, L. Zhou, J. Yang, Z. Ni, Appl. Phys. Lett. 109 (3) (2016)
is only experimentally realized by enhancing chain alignment and 033101.
inter-chain coupling with polymer blends. Furthermore, only the [19] X. Wang, V. Ho, R.A. Segalman, D.G. Cahill, Macromolecules 46 (12) (2013)
4937–4943.
thermal conductivity along the alignment direction is enhanced
[20] B. Zhu, J. Liu, T. Wang, M. Han, S. Valloppilly, S. Xu, X. Wang, ACS Omega 2 (7)
while an isotropic high thermal conductivity material is commonly (2017) 3931–3944.
desirable, which might limit the applications. Blending polymers [21] J. Liu, R. Yang, Phys. Rev. B 81 (17) (2010) 174122.
[22] D.I. Bower, An Introduction to Polymer Physics, Cambridge University Press, New
could only obtain a slightly enhanced thermal conductivity which is York, 2002.
usually lower than 0.5 W·m−1 K−1 and this method is difficult to be [23] M. Wang, S. Lin, Sci. Rep. 5 (2015) 18122.
adopted because of complex synthesis conditions. To further en- [24] M.K. Smith, T.L. Bougher, K. Kalaitzidou, B.A. Cola, MRS Adv. 2 (2017)
3619–3626.
hance the thermal conductivity, appropriate polymer species should [25] Y. Ding, H. Hou, Y. Zhao, Z. Zhu, H. Fong, Prog. Polym. Sci. 61 (2016) 67–103.
be selected based on the theoretical studies before carrying out the [26] J. Ma, Q. Zhang, A. Mayo, Z. Ni, H. Yi, Y. Chen, R. Mu, L.M. Bellan, D. Li,
Nanoscale 7 (40) (2015) 16899–16908.
chain alignment and polymer blending. [27] C. Lu, S.W. Chiang, H. Du, J. Li, L. Gan, X. Zhang, X. Chu, Y. Yao, B. Li, F. Kang,
2) Although incorporating fillers in a polymer could improve the Polymer 115 (2017) 52–59.
thermal conductivity, it is still a challenge to fabricate polymer- [28] C. Canetta, S. Guo, A. Narayanaswamy, Rev. Sci. Instrum. 85 (10) (2014) 104901.
[29] M. Richard-Lacroix, C. Pellerin, Macromolecules 46 (24) (2013) 9473–9493.
based nanocomposites with high thermal conductivity because of [30] A. Arinstein, M. Burman, O. Gendelman, E. Zussman, Nat. Nanotechnol. 2
some well-known difficulties such as filler aggregation and large (2007) 59.
inter-filler thermal contact resistance. 3D fillers, such as graphene [31] Z. Zhong, M.C. Wingert, J. Strzalka, H.H. Wang, T. Sun, J. Wang, R. Chen, Z. Jiang,
Nanoscale 6 (14) (2014) 8283–8291.
foams and carbon foams, should attract more attentions in the fu- [32] H. Sun, J. Phys. Chem. B 102 (1998) 7338–7364.
ture, because of their intrinsic 3D network structure without thermal [33] P.R. Sundararajan, T.A. Kavassalis, J. Chem. Soc., Faraday Trans. 91 (1995)
2541–2549.
contact resistances, which could be more effective than other kinds [34] J. Liu, R. Yang, Phys. Rev. B 86 (10) (2012) 104307.
of fillers, such as 1D CNTs and 2D graphene nanoplates. [35] H. Sun, S.J. Mumby, J.R. Maple, A.T. Hagler, J. Am. Chem. Soc. 116 (1994) 2978.
3) Combining the demonstrated methods emerged for the thermal [36] Y. Xu, X. Wang, J. Zhou, B. Song, Z. Jiang, E.M. Lee, S. Huberman, K.K. Gleason,
G. Chen, Sci. Adv. 4 (3) (2018) eaar3031.
conductivity enhancement in both polymers and polymer nano- [37] T. Zhang, T. Luo, J. Phys. Chem. B 120 (4) (2016) 803–812.
composites together may provide some routes to obtain even higher [38] D.C. Luo, C.L. Huang, Z. Huang, J. Heat. Transf.-ASME 140 (2018) 031302.
[39] H. Ma, Z. Tian, Appl. Phys. Lett. 110 (9) (2017) 091903.
thermal conductivity, for example, mechanically stretching polymer
[40] Z. Guo, D. Le, Y. Liu, F. Sun, A. Sliwinski, H. Gao, P.C. Burns, L. Huang, T. Luo,
blends based nanocomposites filled with graphene foams. Phys. Chem. Chem. Phys. 16 (17) (2014) 7764–7771.
4) To theoretically understand the thermal transport in polymer na- [41] X.P. Chen, Q.H. Liang, J.K. Jiang, C.K.Y. Wong, S.Y.Y. Leung, H.Y. Ye, D.G. Yang,
T.L. Ren, Sci. Rep. 6 (2016) 20621.
nocomposites, many models have been developed to evaluate the [42] X.P. Chen, J.K. Jiang, Q.H. Liang, N. Yang, H.Y. Ye, M. Cai, L. Shen, D.G. Yang,
effect of the size, shape, intrinsic thermal conductivity and disper- T.L. Ren, Sci. Rep. 5 (2015) 16907.
sion of fillers, but few works has explored the effectiveness of hybrid [43] X. Xie, K. Yang, D. Li, T.-H. Tsai, J. Shin, P.V. Braun, D.G. Cahill, Phys. Rev. B 95
(2017) 035406.
fillers [294]. The thermal conductivity mechanisms in composites [44] G. Kikugawa, T.G. Desai, P. Keblinski, T. Ohara, J. Appl. Phys. 114 (3) (2013)
near percolation have yet to be elucidated [295]. 034302.
[45] L. Zhang, M. Ruesch, X. Zhang, Z. Bai, L. Liu, RSC Adv. 5 (107) (2015)
87981–87986.
Acknowledgments [46] E.D. Glowacki, M. Irimia-Vladu, S. Bauer, N.S. Sariciftci, J. Mater. Chem. B 1
(2013) 3742–3753.
[47] T.A. Hubbard, A.J. Brown, I.A.W. Bell, S.L. Cockroft, J. Am. Chem. Soc. 138
CH acknowledges the financial support by the National Natural (2016) 15114–15117.
Science Foundation of China (No. 51406224). RY acknowledges the [48] D. Christendat, T. Abraham, Z. Xu, J. Masliyah, J. Adhes. Sci. Technol. 19 (2005)
financial support by National Science Foundation (Grant No. 0846561, 149–163.
[49] L. Mu, Y. Shi, H. Wang, J. Zhu, ACS Sustain. Chem. Eng. 4 (2016) 1840–1849.
1512776), DOD (Grant No. FA9550-08-1-0078, FA9550-11-1-0109, [50] L. Mu, Y. Shi, X. Guo, J. Wu, T. Ji, L. Chen, X. Feng, X. Lu, J. Hua, J. Zhu, ACS
FA9550-11-C-0034, FA8650-15-1-7524) and ARPA-E (DE-AR DE- Sustain. Chem. Eng. 4 (2016) 3877–3887.
[51] G. Kim, D. Lee, A. Shanker, L. Shao, M.S. Kwon, D. Gidley, J. Kim, K.P. Pipe, Nat.
AR0000743) over the past decade. Mater. 14 (2015) 295–300.
[52] L. Mu, J. He, Y. Li, T. Ji, N. Mehra, Y. Shi, J. Zhu, J. Phys. Chem. C 121 (26) (2017)
References 14204–14212.
[53] G. Gurau, H. Wang, Y. Qiao, X. Lu, S. Zhang, R.D. Rogers, Pure Appl. Chem. 84
(2012) 745–754.
[1] Z.L. Wang, Adv. Mater. 24 (2) (2012) 280–285. [54] J. Liu, R. Yang, Phys. Rev. B: Condens. Matter Mater. Phys. 86 (2012) 104307.
[2] C. Han, Z. Li, S. Dou, Chin. Sci. Bull. 59 (18) (2014) 2073–2091. [55] X. Wei, T. Zhang, T. Luo, Phys. Chem. Chem. Phys. 18 (47) (2016) 32146–32154.
[3] H. Chen, V.V. Ginzburg, J. Yang, Y. Yang, W. Liu, Y. Huang, L. Du, B. Chen, Prog. [56] X. Xie, D. Li, T. Tsai, J. Liu, P.V. Braun, D.G. Cahill, Macromolecules 49 (2016)
Polym. Sci. 59 (2016) 41–85. 972–978.
[4] A.R.J. Hussain, A.A. Alahyari, S.A. Eastman, C. Thibaud-Erkey, S. Johnston, [57] J.E. Mark, Physical Properties of Polymers Handbook, AIP Press, 1996.
M.J. Sobkowicz, Appl. Therm. Eng. 113 (2017) 1118–1127. [58] T. Zhang, X. Wu, T. Luo, J. Phys. Chem. C 118 (36) (2014) 21148–21159.
[5] X. Chen, Y. Su, D. Reay, S. Riffat, Renew. Sustain. Energy Rev. 60 (2016) [59] T. Luo, K. Esfarjani, J. Shiomi, A. Henry, G. Chen, J. Appl. Phys. 109 (7) (2011)
1367–1386. 074321.
[6] H.G. Chae, S. Kumar, Science 319 (5865) (2008) 908–909. [60] R. Tu, Q. Liao, L. Zeng, Z. Liu, W. Liu, Appl. Phys. Lett. 110 (10) (2017) 101905.
[7] S. Shen, A. Henry, J. Tong, R. Zheng, G. Chen, Nat. Nanotechnol. 5 (4) (2010) [61] X. Huang, G. Liu, X. Wang, Adv. Mater. 24 (11) (2012) 1482–1486.
251–255. [62] O. Yamamoto, H. Kambe, Polym. J. 2 (1971) 623–628.
[8] V. Singh, T.L. Bougher, A. Weathers, Y. Cai, K. Bi, M.T. Pettes, S.A. McMenamin, [63] X. Xiong, M. Yang, C. Liu, X. Li, D. Tang, J. Appl. Phys. 122 (3) (2017) 035104.
W. Lv, D.P. Resler, T.R. Gattuso, D.H. Altman, K.H. Sandhage, L. Shi, A. Henry, [64] B. Tonpheng, J. Yu, O. Andersson, Chem. Chem. Phys. 13 (2011) 15047–15054.
B.A. Cola, Nat. Nanotechnol. 9 (5) (2014) 384–390. [65] V. Rashidi, E.J. Coyle, K. Sebeck, J. Kieffer, K.P. Pipe, J. Phys. Chem. B 121 (17)
[9] D. Guo, I. Stolichnov, N. Setter, J. Phys. Chem. B 115 (2011) 13455–13466. (2017) 4600–4609.
[10] X. Zhang, L.J. Richter, D.M. DeLongchamp, R.J. Kline, M.R. Hammond, [66] S. Yu, C. Park, S.M. Hong, C.M. Koo, Thermochim. Acta 583 (2014) 67–71.
I. McCulloch, M. Heeney, R.S. Ashraf, J.N. Smith, T.D. Anthopoulos, B. Schroeder, [67] B. Ni, T. Watanabe, S.R. Phillpot, J. Phys.: Condens. Matter 21 (2009) 084219.
Y.H. Geerts, D.A. Fischer, M.F. Toney, J. Am. Chem. Soc. 133 (2011) [68] C.-W. Nan, R. Birringer, D.R. Clarke, H. Gleiter, J. Appl. Phys. 81 (1997) 6692.
15073–15084. [69] M. Wang, N. Pan, Mater. Sci. Eng. R 63 (2008) 1.
[11] H.C. Yang, T.J. Shin, Z.N. Bao, C.Y. Ryu, J. Polym. Sci. Pt. B-Polym. Phys. 45 [70] K. Pietrak, T.S. Wisniewski, J. Power Technol. 95 (14) (2015).
(2007) 1303–1312. [71] X.J. Wang, L.Z. Zhang, L.X. Pei, J. Appl. Polym. Sci. 131 (8) (2014) 39550.

20
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

[72] A. Boudenne, L. Ibos, M. Fois, J.C. Majeste, E. Gehin, Compos. Part A: Appl. S 36 [131] H. Shin, S. Yang, S. Chang, S. Yu, M. Cho, Polymer 54 (5) (2013) 1543–1554.
(2005) 1545. [132] X. Huang, C. Zhi, P. Jiang, D. Golberg, Y. Bando, T. Tanaka, Adv. Funct. Mater. 23
[73] A.S. Luyt, J.A. Molefi, H. Krump, Polym. Degrad. Stab. 91 (7) (2006) 1629–1636. (14) (2013) 1824–1831.
[74] W. Chen, Z. Wang, C. Zhi, W. Zhang, Compos. Sci. Technol. 130 (2016) 63–69. [133] X. Huang, T. Iizuka, P. Jiang, Y. Ohki, T. Tanaka, J. Phys. Chem. C 116 (25) (2012)
[75] A. Rai, A.L. Moore, Compos. Sci. Technol. 144 (2017) 70–78. 13629–13639.
[76] F. Danes, B. Garnier, T. Dupuis, Int. J. Thermophys. 24 (3) (2003) 771–784. [134] X. Huang, S. Wang, M. Zhu, K. Yang, P. Jiang, Y. Bando, D. Golberg, C. Zhi,
[77] M. Jouni, A. Boudenne, G. Boiteux, V. Massardier, B. Garnier, A. Serghei, Polym. Nanotechnology 26 (1) (2014) 015705.
Compos. 34 (2013) 778. [135] M. Shtein, R. Nadiv, M. Buzaglo, K. Kahil, O. Regev, Chem. Mater. 27 (6) (2015)
[78] R. Kužel, J. Kubát, I. Křivka, J. Prokeš, O. Stefan, C. Klason, Mater. Sci. Eng. B 17 2100–2106.
(1-3) (1993) 190–195. [136] I.L. Ngo, S. Jeon, C. Byon, Int. J. Heat Mass Transfer 98 (2016) 219–226.
[79] J. Kubat, R. Kužel, I. Křivka, P. Bengtsson, J. Prokeš, O. Stefan, Synth. Met. 54 (1- [137] M. TabkhPaz, S. Shajari, M. Mahmoodi, D.Y. Park, H. Suresh, S.S. Park, Compos.
3) (1993) 187–194. Part B-Eng. 100 (2016) 19–30.
[80] A. Boudenne, L. Ibos, E. Géhin, M. Fois, J.C. Majesté, J. Mater. Sci. 40 (16) (2005) [138] A. Suplicz, H. Hargitai, J.G. Kovacs, Int. J. Therm. Sci. 100 (2016) 54–59.
4163–4167. [139] L. Chen, Y.Y. Sun, H.F. Xu, S.J. He, G.S. Wei, X.Z. Du, J. Lin, Compos. Sci. Technol.
[81] I. Krupa, V. Cecen, A. Boudenne, J. Prokeš, I. Novák, Mater. Des. 51 (2013) 122 (2016) 42–49.
620–628. [140] I.E. Afrooz, A. Öchsner, J. Heat Transfer 137 (3) (2015) 034501.
[82] S.H. Lee, Y. Choi, J. Nanosci. Nanotechnol. 13 (2013) 7610–7614. [141] L. Chen, Y.Y. Sun, J. Lin, X.Z. Du, G.S. Wei, S.J. He, S. Nazarenko, Int. J. Heat Mass
[83] Z. Ge, F. Ye, Y. Ding, ChemSusChem 7 (5) (2014) 1318–1325. Transfer 81 (2015) 457–464.
[84] S. Yoshihara, M. Tokita, T. Ezaki, M. Nakamura, M. Sakaguchi, K. Matsumoto, [142] C. Huang, X. Qian, R. Yang, EPL 117 (2017) 24001.
J. Watanabe, J. Appl. Polym. Sci. 131 (6) (2014) 39896. [143] S.H. Xie, Y.Y. Liu, J.Y. Li, Appl. Phys. Lett. 92 (24) (2008) 243121.
[85] L. Sim, S.R. Ramanan, H. Ismail, K.N. Seetharamu, T. Goh, Thermochim. Acta 430 [144] L. Gao, X. Zhou, Y. Ding, Chem. Phys. Lett. 434 (4) (2007) 297–300.
(2005) 155–165. [145] S. Kumar, M.A. Alam, J.Y. Murthy, Appl. Phys. Lett. 90 (10) (2007) 104105.
[86] X.F. Wu, P.K. Jiang, Y. Zhou, J.H. Yu, F.H. Zhang, L.H. Dong, Y. Yin, J. Appl. [146] M. Wang, Q. Kang, N. Pan, Appl. Therm. Eng. 29 (2) (2009) 418–421.
Polym. Sci. 131 (2014) 40528. [147] J.P.M. Péraud, N.G. Hadjiconstantinou, Phys. Rev. B 84 (20) (2011) 205331.
[87] W. Yu, Y. Qi, Y. Zhou, L. Chen, H. Du, H. Xie, J. Appl. Polym. Sci. 133 (13) (2016) [148] W. Tian, R. Yang, Comput. Model. Eng. Sci. 24 (2008) 123.
43242. [149] M. Foygel, R.D. Morris, D. Anez, S. French, V.L. Sobolev, Phys. Rev. B 71 (10)
[88] R. Sun, H. Yao, H.B. Zhang, Y. Li, Y.W. Mai, Z.Z. Yu, Compos. Sci. Technol. 137 (2005) 104201.
(2016) 16–23. [150] Q. Hao, G. Chen, M.S. Jeng, J. Appl. Phys. 106 (11) (2009) 114321.
[89] W.Y. Zhou, C.F. Wang, T. Ai, K. Wu, F.J. Zhao, H.Z. Gu, Compos.: Part A 40 (2009) [151] R. Yang, G. Chen, M.S. Dresselhaus, Phys. Rev. B 72 (12) (2005) 125418.
830–836. [152] I.L. Ngo, C. Byon, Int. J. Heat Mass Transfer 83 (2015) 408–415.
[90] T. Kusunose, T. Yagi, S.H. Firozc, T. Sekinod, J. Mater. Chem. A 1 (2013) [153] I.L. Ngo, S.P. Vattikuti, C. Byon, Int. J. Heat Mass Transfer 102 (2016) 713–722.
3440–3445. [154] F. Zhou, G. Cheng, Comput. Mater. Sci. 92 (2014) 157–165.
[91] W. Zhou, C. Wang, T. Ai, F. Zhao, H. Gu, Compos. Part A 40 (2009) 830–836. [155] M.R. Wang, J.H. He, J.Y. Yu, N. Pan, Int. J. Therm. Sci. 46 (9) (2007) 848–855.
[92] S.Y. Mun, H.M. Lim, H. Ahn, D.J. Lee, Macromol. Res. 22 (2014) 613–617. [156] M. Naffakh, A.M. Díez-Pascual, C. Marco, G. Ellis, J. Mater. Chem. 22 (4) (2012)
[93] W. Zhou, D. Yu, C. Min, Y. Fu, X. Guo, J. Appl. Polym. Sci. 112 (2009) 1695–1703. 1418–1425.
[94] W.Y. Zhou, Q.G. Chen, X.Z. Sui, L.N. Dong, Z.J. Wang, Compos.: Part A 71 (2015) [157] S. Araby, Q. Meng, L. Zhang, H. Kang, P. Majewski, Y. Tang, J. Ma, Polymer 55 (1)
184–191. (2014) 201–210.
[95] Q. Mu, S. Feng, G. Diao, Polym. Compos. 30 (2007) 866–871. [158] M. Russ, S.S. Rahatekar, K. Koziol, B. Farmer, H.X. Peng, Compos. Sci. Technol. 81
[96] J. Zhang, S.H. Qi, Polym. Compos 35 (2014) 381–385. (2013) 42–47.
[97] J.W. Gu, Q.Y. Zhang, J. Dang, J.P. Zhang, Z.Y. Yang, Polym. Eng. Sci. 49 (2009) [159] J. Li, M.L. Sham, J.K. Kim, G. Marom, Compos. Sci. Technol. 67 (2) (2007)
1030–1034. 296–305.
[98] Y. Zhou, Y. Yao, C.Y. Chen, K. Moon, H. Wang, C.P. Wong, Sci. Rep. 4 (2014) 4779. [160] D.P.H. Hasselman, L.F. Johnson, J. Compos. Mater. 21 (1987) 508–515.
[99] T.M.L. Dang, D.H. Yoon, C.Y. Kim, Mater. Chem. Phys. 179 (2016) 204–213. [161] Y. Benveniste, J. Appl. Phys. 61 (1987) 2840–2843.
[100] K. Kim, M. Kim, Y. Hwang, J. Kim, Ceram. Int. 40 (2014) 2047–2056. [162] D.A.G. Bruggeman, Annalen Der Physik 24 (1935) 636–664.
[101] T. Ji, L.Q. Zhang, W.C. Wang, Y. Liu, X.F. Zhang, Y.L. Lu, Polym. Compos. 33 [163] J. Ordonez-Miranda, J.J. Alvarado-Gil, Compos. Sci. Technol. 72 (2012) 853–857.
(2012) 1473–1481. [164] J. Ordonez-Miranda, J.J. Alvarado-Gil, R. Medina-Ezquivel, Int. J. Thermophys. 31
[102] J.W. Gu, Q.Y. Zhang, J. Dang, C. Xie, Polym. Adv. Technol. 23 (2012) 1025–1028. (2010) 975–986.
[103] J.P. Hong, S.W. Yoon, T. Hwang, J.S. Oh, S.C. Hong, Y. Lee, J.D. Nam, [165] J. Ordonez-Miranda, R.G. Yang, J.J. Alvarado-Gil, J. Appl. Phys. 114 (2013)
Thermochim. Acta 537 (2012) 70–75. 064306–064312.
[104] Y. Jiang, Y. Liu, P. Min, G. Sui, Compos. Sci. Technol. 144 (2017) 63–69. [166] A. Patti, P. Russo, D. Acierno, S. Acierno, Compos. Part B-Eng. 94 (2016) 350–359.
[105] H.S. Kim, J.H. Kim, W.Y. Kim, H.S. Lee, S.Y. Kim, M.S. Khil, Carbon 119 (2017) [167] Y. Hwang, M. Kim, J. Kim, Compos. Part A: Appl. Sci. Manuf. 55 (2013) 195–202.
40–46. [168] C.W. Nan, G. Liu, Y. Lin, M. Li, Appl. Phys. Lett. 85 (2004) 3549–3551.
[106] R.R. Nair, P. Blake, A.N. Grigorenko, K.S. Novoselov, T.J. Booth, T. Stauber, [169] A. Norshamira, M. Mariatti, J. Polym. Mater. 33 (1) (2016) 181.
N.M.R. Peres, A.K. Geim, Science 320 (2008) 1308-1308. [170] H. Im, Y. Hwang, J.H. Moon, S.H. Lee, J. Kim, Comp. Part. A: Appl. Sci. Manuf. 54
[107] M. Mehrali, E. Sadeghinezhad, S.T. Latibari, S.N. Kazi, M. Mehrali, (2013) 159–165.
M.N.B.M. Zubir, H.S.C. Metselaar, Nanoscale Res. Lett. 9 (2014) 15. [171] X. Li, X. Fan, Y. Zhu, J. Li, J.M. Adams, S. Shen, H. Li, Comp. Mater. Sci. 63 (2012)
[108] S.H. Song, K.H. Park, B.H. Kim, Y.W. Choi, G.H. Jun, D.J. Lee, B.-S. Kong, K.- 207–213.
W. Paik, S. Jeon, Adv. Mater. 25 (2013) 732–737. [172] E.T. Swartz, R.O. Pohl, Rev. Mod. Phys. 61 (1989) 605.
[109] L.H. Liu, M.D. Yan, J. Mater. Chem. 21 (2011) 3273–3276. [173] M. Hu, P. Keblinski, P.K. Schelling, Phys. Rev. B 79 (2009) 104305.
[110] G. Lin, B.H. Xie, J. Hu, X. Huang, G.J. Zhang, J. Nanomater. 16 (1) (2015) 260. [174] Z. Huang, C. Huang, D. Wu, Z. Rao, Comp. Mater. Sci. 149 (2018) 316–323.
[111] K.M. Shahil, A.A. Balandin, Nano Lett. 12 (2012) 861–867. [175] A. Wurm, D. Lellinger, A.A. Minakov, T. Skipa, P. Pötschke, R. Nicula, I. Alig,
[112] A.A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, C.N. Lau, C. Schick, Polymer 55 (9) (2014) 2220–2232.
Nano Lett. 8 (2008) 902–907. [176] B.P. Grady, J. Polym. Sci. Part B Polym. Phys. 50 (9) (2012) 591–623.
[113] P. Kim, L. Shi, A. Majumdar, P.L. McEuen, Phys. Rev. Lett. 87 (2001) 215502. [177] H. Eslami, M. Behrouz, J. Phys. Chem. C 118 (18) (2014) 9841–9851.
[114] C. Yu, L. Shi, Z. Yao, D. Li, A. Majumdar, Nano Lett. 5 (2005) 1842–1846. [178] M. Mohagheghian, H. Ebadi-Dehaghani, D. Ashouri, S. Mousavian, Compos. Part B
[115] K.I. Winey, T. Kashiwagi, M. Mu, MRS Bull. 32 (4) (2007) 348–353. 42 (2011) 2038.
[116] W. Li, Y. Feng, J. Peng, X. Zhang, J. Heat Transfer 134 (9) (2012) 092401. [179] H. Ebadi-Dehaghani, M. Reiszadeh, A. Chavoshi, M. Nazempour, M.H. Vakili, J.
[117] D.L. Feng, Y.H. Feng, Y. Chen, W. Li, X.X. Zhang, Chin. Phys. B 22 (1) (2013) Macromol. Sci., Part B 53 (1) (2014) 93–107.
016501. [180] M. Bozlar, D. He, J. Bai, Y. Chalopin, N. Mingo, S. Volz, Adv. Mater. 22 (2010)
[118] J. Yang, S. Waltermire, Y. Chen, A.A. Zinn, T.T. Xu, D. Li, Appl. Phys. Lett. 96 (2) 1654–1658.
(2010) 023109. [181] P. Ding, N. Zhuang, X. Cui, L. Shi, N. Song, S. Tang, J. Mater. Chem. C 3 (42)
[119] A.N. Volkov, L.V. Zhigilei, Appl. Phys. Lett. 101 (4) (2012) 043113. (2015) 10990–10997.
[120] L. Kumari, T. Zhang, G.H. Du, W.Z. Li, Q.W. Wang, A. Datye, K.H. Wu, Compos. [182] Y.J. Noh, H.S. Kim, B.C. Ku, M.S. Khil, S.Y. Kim, Adv. Eng. Mater. 18 (2016)
Sci. Technol. 68 (9) (2008) 2178–2183. 1127–1132.
[121] J. Yu, H.K. Choi, H.S. Kim, S.Y. Kim, Compos. Part A: Appl. Sci. Manuf. 88 (2016) [183] H.S. Kim, H.S. Bae, J. Yu, S.Y. Kim, Sci. Rep. 6 (2016) 26825.
79–85. [184] S. Yu, S. Yang, M. Cho, J. Appl. Phys. 114 (21) (2013) 213503.
[122] S.Y. Kim, Y.J. Noh, J. Yu, Compos. Part A: Appl. Sci. Manuf. 69 (2015) 219–225. [185] A. Minnich, G. Chen, Appl. Phys. Lett. 91 (2007) 073105–073107.
[123] I.W. Chiang, B.E. Brinson, R.E. Smalley, J.L. Margrave, R.H. Hauge, J. Phys. Chem. [186] J. Ordonez-Miranda, R.G. Yang, J.J. Alvarado-Gil, Appl. Phys. Lett. 98 (2011)
B 105 (6) (2001) 1157–1161. 233111–233113.
[124] C. Bittencourt, C. Navio, A. Nicolay, B. Ruelle, T. Godfroid, R. Snyders, J.- [187] A. Behrang, M. Grmela, C. Dubois, S. Turenne, P.G. Lafleur, J. Appl. Phys. 114
F. Colomer, M.J. Lagos, X. Ke, G. Van Tendeloo, I. Suarez-Martinez, C.P. Ewels, J. (2013) 014305.
Phys. Chem. C 115 (42) (2011) 20412–20418. [188] A. Behrang, M. Grmela, C. Dubois, S. Turenne, P.G. Lafleur, G. Lebon, Appl. Phys.
[125] S. Huang, L. Dai, J. Phys. Chem. B 106 (14) (2002) 3543–3545. Lett. 104 (2014) 063106.
[126] N. Chopra, M. Majumder, B.J. Hinds, Adv. Funct. Mater. 15 (5) (2005) 858–864. [189] J. Al-Otaibi, G.P. Srivastava, J. Phys.: Condens. Matter 28 (2016) 145304.
[127] J. Zhang, C. Jiang, D. Jiang, H.X. Peng, Phys. Chem. Chem. Phys. 16 (9) (2014) [191] Y. Hu, J. Shen, N. Li, H.W. Ma, M. Shi, B. Yan, W. Huang, W. Wang, M. Ye,
4378–4385. Compos. Sci. Technol. 70 (15) (2010) 2176–2182.
[128] A.J. McNamara, Y. Joshi, Z.M. Zhang, Int. J. Therm. Sci. 62 (2012) 2–11. [192] J. Zhao, F. Du, W. Cui, P. Zhu, X. Zhou, X. Xie, Compos. Part A 58 (2014) 1–6.
[129] K. Gharagozloo-Hubmann, A. Boden, G.J. Czempiel, I. Firkowska, S. Reich, Appl. [193] J. Yang, Y. Yang, S.W. Waltermire, X. Wu, H. Zhang, T. Gutu, Y. Jiang, Y. Chen,
Phys. Lett. 102 (21) (2013) 213103. A.A. Zinn, R. Prasher, T.T. Xu, D. Li, Nat. Nanotechnol. 7 (2) (2012) 91–95.
[130] M. Hu, P. Keblinski, P.K. Schelling, Phys. Rev. B 79 (2009) 104305. [194] Y. Zhou, L. Wang, H. Zhang, Y. Bai, Y. Niu, H. Wang, Appl. Phys. Lett. 101 (1)

21
C. Huang et al. Materials Science & Engineering R 132 (2018) 1–22

(2012) 012903. [244] J.S. Yu, T.E. Lacy Jr., H. Toghiani, C.U. Pittman Jr., Compos. Part B-Eng. 53 (2013)
[195] Z. Han, A. Fina, Prog. Polym. Sci. 36 (7) (2011) 914–944. 267–273.
[196] F. Gong, K. Bui, D.V. Papavassiliou, H.M. Duong, Carbon 78 (2014) 305–316. [245] N. Burger, A. Laachachi, M. Ferriol, M. Lutz, V. Toniazzo, D. Ruch, Prog. Polym.
[197] S.T. Huxtable, D.G. Cahill, S. Shenogin, P. Keblinski, Chem. Phys. Lett. 407 (1) Sci. 61 (2016) 1.
(2005) 129–134. [246] H.L. Zhu, Y.Y. Li, Z.Q. Fang, J.J. Xu, F.Y. Cao, J.Y. Wan, C. Preston, B. Yang,
[198] I.V. Singh, M. Tanaka, M. Endo, Int. J. Therm. Sci. 46 (9) (2007) 842–847. L.B. Hu, ACS Nano 8 (2014) 3606.
[199] S. Shenogin, L. Xue, R. Ozisik, P. Keblinski, D.G. Cahill, J. Appl. Phys. 95 (12) [247] J.D. Renteria, S. Ramirez, H. Malekpour, B. Alonso, A. Centeno, A. Zurutuza,
(2004) 8136–8144. A.I. Cocemasov, D.L. Nika, A.A. Balandin, Adv. Funct. Mater. 25 (2015) 4664.
[200] T.A. El‐Brolossy, S.S. Ibrahim, E.A. Alkhudhayr, Polym. Compos. 36 (7) (2015) [248] Y. Cao, M. Liang, Z. Liu, Y. Wu, X. Xiong, C. Li, X. Wang, N. Jiang, J. Yu, C.T. Lin,
1242–1248. RSC Adv. 6 (72) (2016) 68357–68362.
[201] H. Huang, L. Chen, V. Varshney, A.K. Roy, S. Kumar, J. Appl. Phys. 120 (9) (2016) [249] M.R. Zakaria, M.H.A. Kudus, H.M. Akil, M.Z.M. Thirmizir, Compos. Part B-Eng.
095102. 119 (2017) 57–66.
[202] S. Kaur, N. Raravikar, B.A. Helms, R. Prasher, D.F. Ogletree, Nat. Commun. 5 [250] A.V. Kyrylyuk, P. van der Schoot, Proc. Natl. Acad. Sci. U. S. A. 105 (24) (2008)
(2014) 3082. 8221–8226.
[203] Z. He, X. Zhang, M. Chen, M. Li, Y. Gu, Z. Zhang, Q. Li, J. Appl. Polym. Sci. 129 (6) [251] H.S. Kim, J.U. Jang, J. Yu, S.Y. Kim, Compos. Part B-Eng. 79 (2015) 505–512.
(2013) 3366–3372. [252] S.M. Lebedev, O.S. Gefle, Appl. Therm. Eng. 91 (2015) 875–882.
[204] P.C. Ma, N.A. Siddiqui, G. Marom, J.K. Kim, Compos. Part A: Appl. Sci. Manuf. 41 [253] X.W. Zhao, C.G. Zang, Q.K. Ma, Y.Q. Wen, Q.J. Jiao, J. Mater. Sci. 51 (8) (2016)
(10) (2010) 1345–1367. 4088–4095.
[205] Y. Ni, H. Han, S. Volz, T. Dumitricǎ, J. Phys. Chem. C 119 (22) (2015) [254] S.Y. Kwon, I.M. Kwon, Y.G. Kim, S. Lee, Y.S. Seo, Carbon 55 (2013) 285–290.
12193–12198. [255] J. Li, S. Qi, M. Zhang, Z. Wang, J. Appl. Polym. Sci. 132 (33) (2015) 42306.
[206] Y.D. Kuang, B.L. Huang, Polymer 56 (2015) 563–571. [256] J. Gu, N. Li, L. Tian, Z. Lv, Q. Zhang, RSC Adv. 5 (46) (2015) 36334–36339.
[207] Z. Fan, F. Gong, S.T. Nguyen, H.M. Duong, Carbon 81 (2015) 396–404. [257] W. Li, A. Dichiara, J. Bai, Compos. Sci. Technol. 74 (1) (2013) 221–227.
[208] J. Yu, H. Mo, P. Jiang, Polym. Adv. Technol. 26 (5) (2015) 514–520. [258] X. Zhao, C. Huang, Q. Liu, I.I. Smalyukh, R. Yang, J. Appl. Phys. 123 (2018)
[209] Y. Gao, F. Müller-Plathe, J. Phys. Chem. B 120 (7) (2016) 1336–1346. 085103.
[210] F.P. Du, W. Yang, F. Zhang, C.Y. Tang, S.P. Liu, L. Yin, W.C. Law, ACS Appl. Mater. [259] K.T.S. Kong, M. Mariatti, A.A. Rashid, J.J.C. Busfield, Compos. B. 58 (2014)
Interfaces 7 (26) (2015) 14397–14403. 457–462.
[211] H. Wu, M.R. Kessler, ACS Appl. Mater. Interfaces 7 (10) (2015) 5915–5926. [260] S.Y. Yang, W.N. Lin, Y.L. Huang, H.W. Tien, J.Y. Wang, M.M.C. Ma, Carbon 49 (3)
[212] F. Wang, X. Zeng, Y. Yao, R. Sun, J. Xu, C.P. Wong, Sci. Rep. 6 (2016) 19394. (2011) 793–803.
[213] J.H. Du, J. Bai, H.M. Cheng, Express Polym. Lett. 1 (5) (2007) 253–273. [261] S. Paszkiewicz, A. Szymczyk, R. Pilawka, B. Przybyszewski, A. Czulak,
[214] S. Bal, S.S. Samal, Bull. Mater. Sci. 30 (4) (2007) 379–386. Z. RosŁaniec, Adv. Polym. Techol. 36 (2) (2017) 236–242.
[215] T.C. Clancy, T.S. Gates, Polymer 47 (16) (2006) 5990–5996. [262] F. Zhang, Q. Li, Y. Liu, S. Zhang, C. Wu, W. Guo, J. Therm. Anal. Calorim. 123 (1)
[216] Z. Spitalsky, D. Tasis, K. Papagelis, C. Galiotis, Prog. Polym. Sci. 35 (2010) (2016) 431–437.
357–401. [263] A. Yu, P. Ramesh, X. Sun, E. Bekyarova, M.E. Itkis, R.C. Haddon, Adv. Mater. 20
[217] B. Coto, I. Antia, B. Miren, I. Martinez-de-Arenaza, E. Meaurio, J. Baririga, (2008) 4740.
J.R. Sarasua, Comput. Mater. Sci. 50 (12) (2011) 3417–3424. [264] J.P. Cao, J. Zhao, X. Zhao, F. You, H. Yu, G.H. Hu, Z.M. Dang, Compos. Sci.
[218] W. Yan, Y. Zhang, H. Sun, S. Liu, Z. Chi, X. Chen, J. Xu, J. Mater. Chem. A 2 (48) Technol. 89 (2013) 142–148.
(2014) 20958–20965. [265] H.P. Xu, Z.M. Dang, D.H. Shi, J.B. Bai, J. Mater. Chem. 18 (2008) 2685–2690.
[219] L. Kumari, T. Zhang, G.H. Du, W.Z. Li, Q.W. Wang, A. Datye, K.H. Wu, Ceram. Int. [266] Z. Li, D. Ju, L. Han, L. Dong, Thermochim. Acta 652 (2017) 9–16.
35 (5) (2009) 1775–1781. [267] G. He, X. Zhou, J. Liu, J. Zhang, L. Pan, S. Liu, Polym. Compos. (2017), https://
[220] I. Ahmad, M. Unwin, H. Cao, H. Chen, H. Zhao, A. Kennedy, Y.Q. Zhu, Compos. doi.org/10.1002/pc.24351.
Sci. Technol. 70 (8) (2010) 1199–1206. [268] C. Zhao, S. Xu, Y. Qin, H. Chen, W. Zhao, F. Sun, X. Zhu, Polym. Adv. Technol. 25
[221] C. Chen, H. Wang, Y. Xue, Z. Xue, H. Liu, X. Xie, Y.W. Mai, Compos. Sci. Technol. (12) (2014) 1546–1551.
128 (2016) 207–214. [269] S.Y. Pak, H.M. Kim, S.Y. Kim, J.R. Youn, Carbon 50 (13) (2012) 4830–4838.
[222] D. Ponnamma, K.K. Sadasivuni, Y. Grohens, Q. Guo, S. Thomas, J. Mater. Chem. C [270] Y.J. Xiao, W.Y. Wang, T. Lin, X.J. Chen, Y.T. Zhang, J.H. Yang, Y. Wang,
2 (40) (2014) 8446–8485. Z.W. Zhou, J. Phys. Chem. C 120 (12) (2016) 6344–6355.
[223] C. Zhang, S. Huang, W.W. Tjiu, W. Fan, T. Liu, J. Mater. Chem. 22 (6) (2012) [271] S.Y. Pak, H.M. Kim, S.Y. Kim, J.R. Youn, Carbon 50 (2012) 4830.
2427–2434. [272] K. Kim, J. Kim, Polymer 101 (2016) 168–175.
[224] W.B. Zhang, Z.X. Zhang, J.H. Yang, T. Huang, N. Zhang, X.T. Zheng, Y. Wang, [273] W. Zhang, Z. Zhang, J. Yang, T. Huang, N. Zhang, X.T. Zheng, Y. Wang, Z.W. Zhou,
Z.W. Zhou, Carbon 90 (2015) 242–254. Carbon 90 (2015) 242–254.
[225] H. Ma, L. Tong, Z. Xu, Z. Fang, Nanotechnology 18 (37) (2007) 375602. [274] P.A. Song, L. Liu, S. Fu, Y. Yu, C. Jin, Q. Wu, Y. Zhang, Q. Li, Nanotechnology 24
[226] W.B. Zhang, X.L. Xu, J.H. Yang, T. Huang, N. Zhang, Y. Wang, Z.W. Zhou, Compos. (12) (2013) 125704.
Sci. Technol. 106 (2015) 1–8. [275] K. Wu, C. Lei, R. Huang, W. Yang, S. Chai, C. Geng, F. Chen, Q. Fu, ACS Appl.
[227] X. Zhang, L. Shen, H. Wu, S. Guo, Compos. Sci. Technol. 89 (2013) 24–28. Mater. Interfaces 9 (8) (2017) 7637–7647.
[228] S.S. Rahatekar, K.K. Koziol, S.R. Kline, E.K. Hobbie, J.W. Gilman, A.H. Windle, [276] J. Che, K. Wu, Y. Lin, K. Wang, Q. Fu, Compos. Part A: Appl. Sci. Manuf. 99 (2017)
Adv. Mater. 21 (8) (2009) 874–878. 32–40.
[229] Y.A. Kim, T. Hayashi, M. Endo, Y. Gotoh, N. Wada, J. Seiyama, Scr. Mater. 54 [277] K. Wu, Y. Xue, W. Yang, S. Chai, F. Chen, Q. Fu, Compos. Sci. Technol. 130 (2016)
(2006) 31–35. 28–35.
[230] H. Shen, J. Guo, H. Wang, N. Zhao, J. Xu, ACS Appl. Mater. Interfaces 7 (10) [278] X. Huang, C. Zhi, P. Jiang, J. Phys. Chem. C 116 (44) (2012) 23812–23820.
(2015) 5701–5708. [279] M.T. Pettes, H. Ji, R.S. Ruoff, L. Shi, Nano Lett. 12 (6) (2012) 2959–2964.
[231] H. Huang, C. Liu, Y. Wu, S. Fan, Adv. Mater. 17 (2005) 1652–1656. [280] M. Li, Y. Sun, H. Xiao, X. Hu, Y. Yue, Nanotechnology 26 (10) (2015) 105703.
[232] S. Zhang, Y. Ke, X. Cao, Y. Ma, F. Wang, J. Appl. Polym. Sci. 124 (6) (2012) [281] J. Klett, R. Hardy, E. Romine, C. Walls, T. Burchell, Carbon 38 (7) (2000) 953–973.
4874–4881. [282] R. Sengupta, M. Bhattacharya, S. Bandyopadhyay, A.K. Bhowmick, Prog. Polym.
[233] V. Datsyuk, S. Trotsenko, S. Reich, Carbon 52 (2013) 605–608. Sci. 36 (5) (2011) 638–670.
[234] Z. Lin, Y. Liu, S. Raghavan, K.S. Moon, S.K. Sitaraman, C.P. Wong, ACS Appl. [283] Y.C. Li, G.H. Chen, Polym. Eng. Sci. 47 (882) (2007).
Mater. Interfaces 5 (15) (2013) 7633–7640. [284] J.Q. Wang, W.K. Chow, J. Appl. Polym. Sci. 97 (2005) 366.
[235] H.-B. Cho, T. Nakayama, T. Suzuki, S. Tanaka, W. Jiang, H. Suematsu, K. Niihara, [285] W. Feng, M. Qin, P. Lv, J. Li, Y. Feng, Carbon 77 (2014) 1054–1064.
J. Nanomater. 2011 (2011) 1–7. [286] S. Zhou, L. Yu, X. Song, J. Chang, H. Zou, M. Liang, J. Appl. Polym. Sci. 131 (1)
[236] H. Bai, F. Walsh, B. Gludovatz, B. Delattre, C. Huang, Y. Chen, A.P. Tomsia, (2014) 39596.
R.O. Ritchie, Adv. Mater. 28 (1) (2016) 50–56. [287] S.R. Kim, M. Poostforush, J.H. Kim, S.G. Lee, Express Polym. Lett. 6 (2012) 476.
[237] F. Bouville, E. Maire, S. Meille, B. Van de Moortele, A.J. Stevenson, S. Deville, Nat. [288] H.X. Ji, D.P. Sellan, M.T. Pettes, X.H. Kong, J.Y. Ji, L. Shi, R.S. Ruoff, Energy
Mater. 13 (5) (2014) 508–514. Environ. Sci. 7 (2014) 1185–1192.
[238] S. Deville, E. Saiz, R.K. Nalla, A.P. Tomsia, Science 311 (5760) (2006) 515–518. [289] X. Li, L. Shao, N. Song, L. Shi, P. Ding, Compos. Part A: Appl. Sci. Manuf. 88 (2016)
[239] J. Hu, Y. Huang, Y. Yao, G. Pan, J. Sun, X. Zeng, R. Sun, J.-B. Xu, B. Song, 305–314.
C.P. Wong, ACS Appl. Mater. Interfaces 9 (15) (2017) 13544–13553. [290] X. Zeng, Y. Yao, Z. Gong, F. Wang, R. Sun, J. Xu, C.P. Wong, Small 11 (46) (2015)
[240] D. Stauffer, A. Aharony, Introduction to Percolation Theory, CRC press, 1994. 6205–6213.
[241] J.S. Yu, Jr.T.E. Lacy, H. Toghiani, Jr.C.U. Pittman, Y.K. Hwang, J. Compos. Mater. [291] J. Yin, X. Li, J. Zhou, W. Guo, Nano Lett. 13 (2013) 3232.
45 (2011) 2401–2414. [292] M. Rousseas, A.P. Goldstein, W. Mickelson, M.A. Worsley, L. Woo, A. Zettl, ACS
[242] J.S. Yu, T.E. Lacy Jr., H. Toghiani, C.U. Pittman Jr., J. Compos. Mater. 47 (2013) Nano 7 (2013) 8540.
549–558. [293] J. Chen, X. Huang, Y. Zhu, P. Jiang, Adv. Funct. Mater. 27 (2017) 1604754.
[243] J.S. Yu, T.E. Lacy Jr., H. Toghiani, C.U. Pittman Jr., J. Compos. Mater. 47 (2012) [294] A. Agrawal, A. Satapathy, Int. J. Therm. Sci. 89 (2015) 203–209.
1273–1282. [295] C.W. Nan, Y. Shen, J. Ma, Annu. Rev. Mater. Res. 40 (2010) 131–151.

22

You might also like