You are on page 1of 7

International Journal of Heat and Mass Transfer 100 (2016) 355–361

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

A numerical investigation on the heat conduction in high filler loading


particulate composites
Zhen Tong a, Meng Liu b, Hua Bao a,⇑
a
University of Michigan-Shanghai Jiao Tong University Joint Institute, Shanghai Jiao Tong University, Shanghai 200240, China
b
School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China

a r t i c l e i n f o a b s t r a c t

Article history: Particle-filled composite materials have been widely used as thermal interface materials (TIMs) to reduce
Received 3 February 2016 the thermal contact resistance. For industrial applications, the particle-filled composite usually has high
Accepted 25 April 2016 volume fraction (>50%). However, most of the research on the thermal properties of particle-filled com-
Available online 10 May 2016
posites has been focusing on low volume fraction composites. In this work, the finite element method
(FEM) is adopted to investigate the particle-filled composites with high filler loading. We consider the
Keywords: close-packed simple cubic (SC), face-centered cubic (FCC), and a dual diameter (DD) model with even
Finite element method
a higher volume fraction than the FCC. It is found that with a high volume fraction, small increase in vol-
Composite material
Thermal conductivity
ume fraction can lead to a strong enhancement in the overall thermal conductivity. With a certain filler
High volume fraction loading and thermal conductivity of the matrix, the effective thermal conductivity first dramatically
increases with the thermal conductivity of the filler and then saturates. We show that the effective med-
ium theory based models cannot properly predict the effective thermal conductivity for the close-packed
structures. The percolation theory based on the resistance network agrees surprisingly well with our sim-
ulation results. Through a careful investigation of the effect of proximity between adjacent particles, it is
found that good contact between particles is crucial to the enhancement of the overall thermal conduc-
tivity. We also considered the interface thermal resistance between fillers and matrix and compared the
simulation results with analytical models. Our analysis provides a better understanding on the heat
transfer in the high volume fraction composite materials and is important for the fabrication of high ther-
mal conductivity TIMs.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction Recent experimental investigations of particle filled TIMs


mainly take two different approaches, i.e., increasing the volume
With the increasing power density of electronic device, heat fraction and mixing different types of fillers together. For example,
dissipation has become one of the most critical challenges [1]. TIMs Yu et al. [5] reported that the composite of hybrid size alumina fil-
[2] have been widely used in electronic packaging and cooling sys- lers with the addition of graphene of 1 wt.% filled in silicone matrix
tem to reduce the contact thermal resistance across jointed sur- had an effective thermal conductivity of 3.45 W/m K at a filler vol-
faces, such as the contact between microprocessors and heat ume fraction of 63%. Gao et al. [6] found that the alumina with a
sinks [3]. Particle-filled composite materials made of thermal diameter of 75 lm filled silicone matrix composite had an effective
grease matrix (with typical thermal conductivity of 0.1–0.4 W/ thermal conductivity of 2.25 W/m K at a high volume fraction of
m K [3,4]) and highly conductive ceramic or metallic particles are 62%. Some commercial TIMs even report thermal conductivity as
the most widely used commercial TIMs, of which the thermal con- high as 17 W/m K [7], although the data have not been verified
ductivity values are generally in the range of 2–5 W/m K [4]. by other sources. It can be seen that the reported effective thermal
Although there are practical concerns such as thermal stability, conductivity of composites with high volume fraction varies signif-
wetting ability, and viscosity, to further increase the effective ther- icantly in different studies.
mal conductivity of TIMs is still highly desirable to the electronic Actually, the effective thermal conductivity of composites is
packaging industry. determined by many factors [8] other than the volume fraction,
such as the particle type and shape, the collective arrangement
of particles, and the interfacial thermal resistance between fillers
⇑ Corresponding author. Tel.: +86 21 34205660.
and matrix. It is generally difficult to separate the effect of these
E-mail address: hua.bao@sjtu.edu.cn (H. Bao).

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.04.092
0017-9310/Ó 2016 Elsevier Ltd. All rights reserved.
356 Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361

Nomenclature

T1, T2 (K) temperature d (m) thickness of the interfacial layer


kf (W/m K) thermal conductivity of filler kc (W/m K) thermal conductivity of the cushion
km (W/m K) thermal conductivity of matrix ks (W/m K) thermal conductivity of the shell
keff (W/m K) effective thermal conductivity Rb (m2 K/W) interfacial thermal resistance
Q (W) heat flow rate Rs (m2 K/W) thermal resistance of the shell
A (m2) cross-sectional area Rm (m2 K/W) the matrix intrinsic thermal resistance correspond-
DT (K) temperature difference ing to thickness of d
L (m) length of the cubic model Bi Boit number
D (m) diameter of the particle f2 geometric parameter
d (m) gap distance b1 dimensionless parameter
f volume fraction R dimensionless parameter
DS (m2) contact area of the cushion k ratio of thermal conductivity of fillers to matrix
S (m2) surface area of the particle a dimensionless parameter
t (m) the half of the thickness of the cushion d0 (m) critical diameter

factors by purely experimental approach. Therefore, theoretical arrangement of particles. Numerical simulations were performed
understanding is vital to guide the design and fabrication of better to study different types of fillers and different arrangements of fil-
composite material for thermal application. There are two main lers, such as cubic [25], spherical elliptic fillers [26], and fibers [27],
categories of analytical or semi-analytical theories. One is the with ordered or random distribution [28]. The lattice Boltzmann
effective medium theory. The widely used models include the method is also employed to calculate the effective thermal conduc-
Maxwell [9] and Bruggeman [10] symmetric model. These simple tivity of particulate TIMs in squeeze flow [29] and study the heat
models only consider the effect of the volume fraction and are gen- transfer in porous media [30]. Most of these studies are focusing
erally valid for low volume fraction (usually less than 40 vol% [11]). on composites with relatively low filler volume fraction as well.
Another approach is based on the Rayleigh’s mathematical tech- There are also some methods developed for high volume fraction,
nique of expressing the temperature field as a multipole expansion such as the Agari’s [31] semi-empirical model which requires
in the composites [12]. This technique works for cubic arrange- experimental data to obtain the fitting curve as the function of fil-
ments (including SC, BCC, FCC) of spherical fillers and gives more ler concentration. Some other methods are based on the assump-
accurate solutions for larger volume fraction. However, the Ray- tion that the composites can be treated as resistance networks.
leigh’s original treatment was later questioned by Levine et al. For example, the effective unit cell model (EUCM) [32] uses the
[13] because it just considered low order multipole expansion effective medium theory to determine the value of the resistance,
coefficients and involved the summation of a non-absolutely con- and another model uses the FEM [33] to obtain the value of the
vergent series. In order to overcome these difficulties, McPhedren resistance. The percolation model proposed by Devpura et al.
et al. [14] and McKenzie et al. [15] extended the Rayleigh’s work [34] models the composite material as a binary compound of cubic
to take into account multipoles of arbitrarily high order and over- units with the same size and then uses the resistance network to
come the non-absolutely convergent difficulty for the case of SC obtain the effective thermal conductivity of the composite. It needs
particle arrangement and the case of BCC and FCC, respectively. an empirical correction factor to be applied to spherical fillers.
Sangani et al. [16] modified the multipole expansion by taking These models for a high volume fraction either need fitting param-
advantages of the symmetric particles and the periodicity of the eters, or the assumption that the system can be modeled as a resis-
temperature to calculate the effective thermal conductivity for tance network. In contrast, a direct numerical simulation of the
cubic array of spheres (including SC, BCC, FCC) from dilute to high filler loading composite materials does not require any
nearly touching volume fraction. There are also other approaches assumptions. It can help to develop a better understanding of the
that investigate the effective thermal conductivity of composites heat transfer mechanism in the high filler loading composite and
for cubic array of spheres, such as the boundary integral method also to validate the numerous models that have been proposed in
developed by Zick [17] and the asymptotic expression derived by literature.
Keller [18]. However, these works generally only consider cubic In this work, a numerical simulation of particle-filled compos-
lattice arrangement of particles and are difficult to apply to volume ites is carried out, focusing on high filler loading composites model.
fraction higher than the closed-packed FCC structure. To further A few high volume fraction unit cell models are established. On the
include the effect of interface thermal resistance, several exten- premise of neglecting the microscale effect, the FEM is employed to
sions of the analytical theories have been developed. For example, investigate the effects on the effective thermal conductivity of
Hasselman and Johnson [19] derived an expression based on Max- these cells. The factors, such as volume fraction, thermal conduc-
well model that includes the effect of interface thermal resistance. tivity of the filler, the proximity effect of closely packed fillers,
With the modification of Bruggeman model, Every et al. [20] pro- and the interfacial thermal resistance between fillers and matrix,
posed an expression that includes a dimensionless parameter Biot have been analyzed. The results are also compared with different
number to consider the interfacial thermal resistance between fil- models and discussed.
lers and matrix. Cheng et al. [21] applied the Rayleigh’s method to
calculate the effective thermal conductivity of periodic arrays of 2. Model description and simulation details
spherical inclusions with interfacial thermal resistance, and an
accurate approximation formula was provided. Some of the models To model high volume fraction composite materials, ideally one
are summarized in review articles [22,23]. should build a large simulation domain and have many particles
On the other hand, the direct numerical solution of the heat randomly placed in the domain [35]. However, for the high volume
diffusion equation [24] can obtain detailed information of heat fraction configurations, many particles will be close to each other.
transfer in composites and theoretically consider any type of As will be shown later, the proximity effect between adjacent
Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361 357

particles has a strong effect on the overall thermal conductivity, we see that our result agrees well with the Sangani’s result when
and extremely fine mesh has to be used to properly resolve the the ratio of gap to the diameter (d/D) is larger than 0.001. It starts
gap between adjacent particles. Therefore, to achieve a high vol- to deviate from the semi-analytical value when d/D is smaller than
ume fraction, keep a reasonable amount of computational cost, 0.001. This indicates that our mesh can properly resolve the gap as
and preserve the accuracy, we considered ordered arrangement long as d/D is larger than 0.001. Therefore, in all the subsequent
of close-packed fillers in the composite. Three composite models simulations, the models are built so that d/D is 0.001. For the DD
have been established, including simple cubic (SC), face-centered case, the gaps between any two large particles are kept the same
cubic (FCC) and a dual diameter (DD) model by filling the voids value (d/D = 0.001). It should be pointed out that when the gap
in FCC model with other smaller size particles. The unit cells of between the spheres is vanishing, the effective thermal conductiv-
these models are shown in Fig. 1. In the subsequent discussions, ity of the composite material increases significantly, as shown in
we define D to be the diameter of the particle in the SC or FCC Fig. 2(b). Such a conclusion is true as long as the particles have
model, and the diameter of larger particle in the DD model. much larger thermal conductivities than the matrix material
In the finite element simulation, a representative cubic unit cell [14,15], which is usually the case for TIMs. Therefore, one should
is used. We take the SC unit cell model shown in Fig. 1(a) as an take extra caution while performing FEM simulation for those
example. The upper and lower surface (i.e., ABCD and EFGH) are nearly touching spheres. Very careful convergence tests have to
set to be constant temperatures T1 and T2, respectively. The other be carried out before producing any useful data. This is also partly
four surfaces are assumed to be adiabatic. For such a model, the the reason why we are not able to build a much larger domain and
effective thermal conductivity of the unit cell can be calculated by perform simulations of pseudorandom configurations with a large
number of particles.
QL
keff ¼ ; ð1Þ
ADT
where keff is the effective thermal conductivity, Q is the heat flow 3.2. Effect of volume fraction and thermal conductivity of fillers
rate, A is the cross-sectional area of the heat flow, DT is the temper-
ature difference between upper and lower surface (T 1  T 2 ), and L is The effective thermal conductivity values of composites are cal-
the height of the simulation domain. The tetrahedral element is culated for different models, as shown in Fig. 3. It should be noted
used to mesh the model, and then the steady state heat diffusion that d is 0.001D for SC and FCC models, which gives 52.20% and
equation is solved by FEM. The heat flow rate Q, across the upper 73.83% volume fraction, respectively. For the DD model, gap dis-
and lower surface, is calculated by integrating the heat flux on tance between large spheres is 0.001D, while the gap distance
the corresponding surface. between small sphere and large sphere may vary (so is the volume
fraction). The 79.05% volume fraction case corresponds to the case
3. Results and discussion in which the gap between a small sphere and a large sphere is
0.001D. In this calculation, we consider how the effective thermal
3.1. Convergence test conductivity varies with the thermal conductivity ratio of filler to
matrix (kf =km ). It should be noted that for a given structure the
Unlike the low volume fraction case, for close-packed struc- ratio of the effective thermal conductivity to the matrix thermal
tures, the fillers are treated as spheres that are tangential to each conductivity (keff =km ) is only a function of kf =km [37]. From Fig. 3,
other. However, an exact tangential structure has infinitely small we can see that keff =km first significantly increases with the
feature size, which cannot be resolved with a practical mesh den- increase of kf =km and then saturates at around 1000 (the relative
sity. A small gap d always exists between neighboring particles, as increment is within 2% with even larger kf =km ). It shows that it is
shown in the inset of Fig. 2(a). In our simulations, the smallest ele- unnecessary to seek for fillers with extremely high thermal con-
ment has an edge length of 105 of the side length of the cube and ductivity, since the enhancement of the effective thermal conduc-
the number of elements varies from 3  105 to 2  106 depending tivity is limited. On the other hand, if the thermal conductivity of
on the particular model. To find out the smallest gap that our mesh the low conductivity matrix can be enhanced, the effective thermal
can resolve, we tested different gap distances and compared our conductivity will be enhanced proportionally. The different depen-
FEM results with the semi-analytical data provided by Sangani dence on the matrix and filler is due to the fact that the matrix is
et al. [16]. Since the data for FCC and infinite particle thermal con- thermally connected while the fillers are not [38]. If further com-
ductivity were available in these references [16,36], this case was paring the thermal conductivity of three different structures with
chosen to validate our simulation results. The thermal conductivity different volume fractions, we find that the larger volume fraction
ratio of the particle to the matrix is set to be 107 in our simulations always gives higher effective thermal conductivity. In addition, we
to approximate the infinite thermal conductivity. From Fig. 2(a), also calculate the effective thermal conductivity of three DD

Fig. 1. Composite unit cell models. (a) SC composite unit cell model, the km represents the matrix, and the kf represents the filler, the same as below. (b) FCC composite unit
cell model. (c) DD composite unit cell model. The smaller particles have the same size.
358 Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361

Fig. 2. (a) The ratio of effective thermal conductivity to matrix thermal conductivity (keff/km) varies with the ratio of the gap distance to the diameter of the sphere (d/D) for a
FCC structure with sphere to matrix thermal conductivity ratio of infinity using our method (square dot) and the analytical results (red curve [16,36]). The top view of FCC
structure is shown in the inset, where D is the diameter of the sphere, L is the lateral length of the cube, and d is the gap distance. (b) The ratio of effective thermal
conductivity to matrix thermal conductivity (keff/km) varies
pffiffi
with the volume fraction f for a FCC structure using our method (circular dot) and the analytical results (red curve
[16,36]). The vertical dashed line corresponds to f ¼ 62 p. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)

are shown in Table 1. We can see that the result of Maxwell model
is nearly 50% less than that of present work and the Bruggeman
symmetric model gives nearly 5 times larger value than the simu-
lation results. Since these models are just derived at relatively low
volume fractions, it is easy to understand that they cannot work for
the high loading particle-filled composites.
We also compare our results with the percolation model devel-
oped by Devpura et al. [34]. Following their method we developed
a MATLAB code to implement the transfer matrix method. As this
method works only for cubic fillers, following the suggestion by
Devpura et al., the volume fraction was multiplied by the correc-
tion factor of 0.637 to produce the results for spherical particles.
The thermal conductivity values of the two components are also
30 W/m K and 0.2 W/m K, to model the alumina–silicone compos-
ites. The corresponding width and height of the system were con-
sidered to be 25, the length was to be 800. To reduce the statistical
uncertainty, for each volume fraction, the final value is achieved by
Fig. 3. In the left-bottom axis, these curves describe keff/km as a function of kf/km. In
averaging the calculated value of 20 independent systems. Surpris-
the right-top axis, these curves describe the effective thermal conductivity as a
function of the thermal conductivity of filler at a given matrix (km = 0.2 W/m K).
ingly, the percolation model gives relatively smaller effective ther-
mal conductivity than our simulation results, but the difference is
within 10%. This actually validates their percolation model to pre-
models with different volume fractions by changing the size of the dict the thermal conductivity of high filler loading composite
small sphere. From Fig. 3, we find that the increment of the effec- materials.
tive thermal conductivity of DD model with low volume fraction is
not obvious when compared to that of FCC model, but it is more 3.3. Effect of the contact area
significant with higher volume fraction corresponding that the
smaller sphere is nearly in contact with the large sphere. In the previous discussions, the particles in these high filler
We further consider the widely used alumina–silicone compos- loading composite models are assumed to be perfectly spherical
ite, in which the thermal conductivity of alumina and the silicone and nearly tangential to each other. A valid question would be
can be assumed to be 30 W/m K and 0.2 W/m K, respectively. With what happens if the two neighboring spheres are exactly tangen-
the thermal conductivity ratio kf =km to be 150, the effective ther- tial to each other. However, such a problem is ill defined in the
mal conductivity values of the alumina–silicone composites can framework of FEM because it has infinitely small feature size.
be determined from Fig. 3, which are 1.71 W/m K for SC composite Moreover, in reality such a model may not even be necessary to
model, 3.65 W/m K for FCC composite model, and 4.36 W/m K for describe the contact of two particles because the surface of the par-
DD composite model. Compared to FCC model, the relative incre- ticles more or less has certain roughness. Based on these consider-
ment of the volume fraction of the DD model is just 7%, but the ations, we replace the contact point for the tangential spheres with
increment of the effective thermal conductivity is nearly 20%. It a cylindrical ‘‘cushion”, as shown in Fig. 4(a). Note that while build-
is shown that the effective thermal conductivity of composites ing these models the two spheres are kept perfectly tangential to
with high filler volume fraction is very sensitive to the variation each other, and we just replace the region close to the tangential
of the filler loading. Even for a small amount of extra fillers, the point with a cylindrical cushion. Therefore, the half of the thickness
enhancement of thermal conductivity is significant. of the cushion t and its area DS is related by
In addition, we compare our results with the theoretical models
in the case of alumina–silicone composites, and the comparisons DS ¼ ptðD  tÞ; ð2Þ
Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361 359

Table 1
Comparisons between our simulation results and the theoretical models. Here we assume kf = 30 W/m K and km = 0.2 W/m K.

Model Expressions SC (W/m K) FCC (W/m K) DD (W/m K)


Vol (%) – 52.20 73.83 79.05
Maxwell keff kf þ2km þ2f ðkf km Þ 0.83 1.77 2.27
km
¼ kf þ2km f ðkf km Þ
Bruggeman symmetric k k kf keff 8.93 18.42 20.75
ð1  f Þ kmmþ2keff þ f kf þ2keff
¼0
eff

Percolation – 1.61 ± 0.15 3.32 ± 0.14 4.28 ± 0.13


FEM – 1.71 3.65 4.36

km is the thermal conductivity of matrix, kf is the thermal conductivity of filler, f is the volume fraction of fillers.

Fig. 4. (a) The cylindrical ‘cushion’ model. Here kc is the thermal conductivity of the cushion, t is half of the thickness of the cushion, and DS is the area of the cushion (contact
area). (b) The effective thermal conductivity varies with the ratio of the contact area to the surface area of the particle (S is the surface area of the particle).

where D is the diameter of the particle. Due to the complexity of conductivity. Although it is previously known that forming con-
building such models, we only consider the SC arrangement of par- ductive pathway is important to make high conductivity composite
ticles to understand the importance of the contact. It should be material, this simulation results show that the detail at the contact
noted that the thermal conductivity of the thin layer cannot be between particles is also a key factor to obtain high effective ther-
determined directly because it is related to too many factors, mal conductivity of composites.
including the surface roughness, pressure, and the complicated
micro/nanoscale thermal transport effects [39]. Nevertheless, we 3.4. Effect of the interfacial thermal resistance
can estimate the upper and lower limit of the thermal conductivity
of the thin layer, which are the thermal conductivity of filler and In the simulation results presented above, the effect of interface
matrix, respectively. The simulation is performed assuming the fil- thermal resistance has not been considered. The results should be
ler and matrix are alumina and silicone, as mentioned above. The useful if the particles are relatively large so that such effect can be
effective thermal conductivity is then calculated with different con- neglected. Sometimes it is practically needed to add smaller parti-
tact areas and the thermal conductivity values of the cushion are cles into the TIMs. Therefore, it is also necessary to quantify the
chosen to be 0.2, 1.0, 3.0, and 30.0 W/m K. When the thermal con- effect of interface thermal resistance for high volume fraction
ductivity of the cushion is 0.2 W/m K, the particles are separated particle-filled composite. We employed a core–shell model as
by a layer of matrix material, corresponding to a poor contact. If shown in the inset of Fig. 5. The interfacial thermal resistance
it is 30.0 W/m K, this is an ideal contact (particles are perfectly was modeled as a very thin layer between fillers and matrix (i.e.,
connected). d=D ¼ 0:01 [26] in the present work), and the value of the interfa-
The simulation results are shown in Fig. 4(b), in which the cial thermal resistance was incorporated in the dimensionless
lower horizontal axis is the ratio of the cushion area to the particle parameter Biot number
surface area, and the upper horizontal axis is the ratio of the thick-
ness of the corresponding thin layer to the diameter of the particle. Rb km
Bi ¼ ; ð3Þ
From Fig. 4(b), we can see that the effect of the contact area D
behaves completely differently for cushions of different thermal where Rb is the interfacial thermal resistance, km is the thermal con-
conductivities. When the thermal conductivity of the cushion is ductivity of the matrix, D is the diameter of fillers. The critical diam-
larger than 3.0 W/m K, the effective thermal conductivity of the eter d0 (=2Rbkm) is defined by setting the Biot number equal to 1
composite significantly increases with contact area. On the other [40]. The critical diameter has clear physical meaning. If the particle
hand, if the cushion thermal conductivity is below 1.0 W/m K, diameter is less than the critical diameter, the thermal conductivity
the effective thermal conductivity decreases with contact area. of the composite will be less than that of the matrix and it will be
The trend is depending on whether a conductive pathway can be greater than that of matrix otherwise. Due to the complexity of such
formed in the structure. It can also be noticed that when the cush- structures, we did not consider the DD model here. To ensure the
ion thermal conductivity is larger than 3.0 W/m K, even a very accuracy, the gap between the outer shells of neighboring spheres
small contact area can significantly enhance the overall thermal is kept 0.001D. Due to the thin shell layer, the volume fractions of
360 Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361

sely proportional to the particle size, so keff/km decreases with the


decrease of the particle size. This is because when the particle size
is smaller, the interfaces become denser and the interface thermal
resistance becomes dominant. As such, for a fixed interfacial ther-
mal resistance, there is a critical diameter, denoted by d0 in Fig. 5.
The particles can increase the thermal conductivity of the compos-
ite when the diameter of the particle is larger than d0, and it will
decrease the thermal conductivity otherwise. We notice that the
results calculated by the formula provided by Cheng et al. agree
well with the FEM results. However, there exists a discrepancy
between Maxwell–Garnet and the FEM results (and Cheng’s for-
mula), especially for small Bi. We believe this is still mainly due
to the dipole approximation of the Maxwell–Garnet model and it
is typically valid for low volume fraction (less than 40% [4]). For
larger Bi, since the interface resistance becomes dominated, the
results from different models agree better with each other.
As for alumina–silicone composite, the value of the interfacial
Fig. 5. The ratio of the effective thermal conductivity to the matrix thermal
thermal resistance has not been directly measured. We then take
conductivity (keff/km) varies with Biot number, calculated by FEM (dots), Cheng et al. the value to be (3.75 ± 1.25 ) 108 m2 K/W, based on the mea-
(solid curve [21]), and Maxwell–Garnet model (dashed curve [20]). The inset is the surement of Putnam et al. [42] for the interface thermal resistance
FCC core–shell model used to investigate the effect of interfacial thermal resistance. between alumina and PMMA. It is assumed here that the interfacial
A thin shell is used to model the interfacial thermal resistance between fillers and
thermal resistance between alumina and polymers (i.e., silicone
matrix.
and PMMA) does not differ significantly. Then we can determine
that the critical diameter of alumina for alumina–silicone compos-
ite is 15 nm, and the diameter corresponding to the effective ther-
fillers in these models are slightly smaller than those in Section 3.2. mal conductivity is nearly independent (when Bi is less than 103)
The values of the ratio of effective thermal conductivity to matrix of the Biot number is 7.5 lm. Therefore, we can see that the effec-
thermal conductivity (keff/km) are plotted in Fig. 5 as a function of tive thermal conductivity strongly depends on the particle diame-
Biot number. In the same figure, for comparison purpose, we also ter when 15 nm < D < 7:5 lm, and it becomes lower than the
show the values calculated by using the formula derived by Cheng thermal conductivity of the matrix when D < 15 nm. Our simula-
et al. [21] (based on Rayleigh’s multipole expansion method) for tion result can be a good guidance to the choice of the size of the
cubic arrays of spheres with interface thermal resistance. Based alumina fillers to fabricate high thermal conductivity TIMs.
on this model, the effective thermal conductivity is given by

keff 1 þ 2f b1  2ð1  f Þf2 b21 4. Conclusions


¼ ; ð4Þ
km 1  f b1  2ð1  f Þf2 b21
To summarize, FEM simulations on the heat conduction of
where keff is the effective thermal conductivity, km is the thermal close-packed particulate composite materials are performed. The
conductivity of matrix, f is the volume fraction of fillers, f2 is the effects of thermal conductivity of fillers, volume fraction, the prox-
geometric parameter which has been tabulated as a function of imity effect of adjacent particles, and interfacial thermal resistance,
the volume fraction by McPhedran et al. [41]. By definition the on the overall thermal conductivity of the composite, were inves-
dimensionless parameter b1 ¼ kþ2þ2R k1R
where k ¼ kf =km and tigated. A few conclusions can be drawn. First, the effective ther-
R ¼ 2Rb kf =D ¼ 2kBi. In addition, we also provide the values calcu- mal conductivity of the composite first increases with the
lated by using the Maxwell–Garnet model [20] with the effective thermal conductivity of fillers and then saturates when the filler
thermal conductivity given by, thermal conductivity is 1000 times larger than the matrix. Second,
the effective thermal conductivity of composites with high filler
keff ½kf ð1 þ 2aÞ þ 2km  þ 2f ½kf ð1  aÞ  km  volume fraction is sensitive to the variation of the filler loading
¼ ; ð5Þ
km ½kf ð1 þ 2aÞ þ 2km   f ½kf ð1  aÞ  km  especially at the nearly contact state. Even for small enhancement
where km is the thermal conductivity of matrix, kf is the thermal of filler loading, the enhancement of thermal conductivity is signif-
conductivity of filler, and f is the volume fraction of fillers. By defi- icant. Third, details at the contact region between adjacent parti-
cles are crucial to the effective thermal conductivity of
nition the dimensionless parameter a ¼ 2Rb km =D ¼ 2Bi.
composites. More effective thermal contacts between particles
In our simulations, the thermal resistance of the shell
can help to build the conductive pathway and significantly
Rs ¼ Rb þ Rm ¼ d=ks , where Rm is the matrix intrinsic thermal resis-
enhance the overall thermal conductivity. Fourth, the interfacial
tance corresponding to the thickness of d, and ks is the thermal
thermal resistance and particle size is also an important factor
conductivity of the shell. The effect of the interface thermal resis-
for Biot number larger than 0.001. In the case of alumina–silicone
tance on the overall thermal conductivity is shown in Fig. 5, from
composites, interface thermal resistance for particles larger than a
which we can see that the interfacial thermal resistance has a sig-
few microns is generally negligible. The particles with diameter
nificant effect for high filler loading composites. Furthermore, one
smaller than 15 nm will have negative contributions to the thermal
can see that the variation of keff/km is within 2.5% when Bi is less
conductivity.
than 103 but significantly decreases in the range of 103 to 0.5,
and then decreases to less than 1 when Bi is less than 0.5. This phe-
nomenon can be understood from the following two aspects. For Acknowledgement
fixed particle size, the Biot number is proportional to the interfacial
thermal resistance. Then keff/km decreases with the increase of The authors gratefully acknowledge the support from HUAWEI
interfacial thermal resistance, as it induces additional resistance. TECHNOLOGIES CO., LTD. Fruitful discussions with Yan Xu and
For a fixed interfacial thermal resistance, the Biot number is inver- Wenbin Liu are gratefully appreciated.
Z. Tong et al. / International Journal of Heat and Mass Transfer 100 (2016) 355–361 361

References [22] P. Cheng, C.-T. Hsu, The effective stagnant thermal conductivity of porous
media with periodic structures, J. Porous Media 2 (1) (1999) 2.
[23] J. Ma, C.W. Nan, Effective-medium approach to thermal conductivity of
[1] E. Pop, Energy dissipation and transport in nanoscale devices, Nano Res. 3 (3)
heterogeneous materials, Ann. Rev. Heat Transfer 17 (2014) 303–331.
(2010) 147–169.
[24] K. Ramani, A. Vaidyanathan, Finite element analysis of effective thermal
[2] D.D.L. Chung, Thermal interface materials, J. Mater. Eng. Perform. 10 (1) (2001)
conductivity of filled polymeric composites, J. Compos. Mater. 29 (13) (1995)
56–59.
1725–1740.
[3] R. Prasher, Thermal interface materials: historical perspective, status, and
[25] I.A. Tsekmes, R. Kochetov, P.H.F. Morshuis, J.J. Smit, Modeling the thermal
future directions, Proc. IEEE 94 (8) (2006) 1571–1586.
conductivity of polymeric composites based on experimental observations,
[4] R.J. Linderman, T. Brunschwiler, U. Kloter, H. Toy, B. Michel, Hierarchical nested
IEEE Trans. Dielectr. Electr. Insul. 21 (2) (2014) 412–423.
surface channels for reduced particle stacking and low-resistance thermal
[26] J. Zeng, R. Fu, S. Agathopoulos, S. Zhang, X. Song, H. He, Numerical simulation
interfaces, in: Twenty Third Annual IEEE Semiconductor Thermal
of thermal conductivity of particle filled epoxy composites, J. Electron. Packag.
Measurement and Management Symposium, SEMI-THERM 2007, 2007, pp.
131 (4) (2009) 041006.
87–94.
[27] S.-T. Tu, W.-Z. Cai, Y. Yin, X. Ling, Numerical simulation of saturation behavior
[5] W. Yu, H. Xie, L. Yin, J. Zhao, L. Xia, L. Chen, Exceptionally high thermal
of physical properties in composites with randomly distributed second-phase,
conductivity of thermal grease: synergistic effects of graphene and alumina,
J. Compos. Mater. 39 (7) (2005) 617–631.
Int. J. Therm. Sci. 91 (2015) 76–82.
[28] L.C. Davis, B.E. Artz, Thermal conductivity of metal-matrix composites, J. Appl.
[6] B.Z. Gao, J.Z. Xu, J.J. Peng, F.Y. Kang, H.D. Du, J. Li, S.W. Chiang, C.J. Xu, N. Hu, X.S.
Phys. 77 (10) (1995) 4954–4960.
Ning, Experimental and theoretical studies of effective thermal conductivity of
[29] R.H. Khiabani, Y. Joshi, C.K. Aidun, Thermal properties of particulate TIMs in
composites made of silicone rubber and Al2O3 particles, Thermochim. Acta 614
squeeze flow, Int. J. Heat Mass Transfer 53 (19) (2010) 4039–4046.
(2015) 1–8.
[30] M. Wang, N. Pan, J. Wang, S. Chen, Mesoscopic simulations of phase
[7] Thermal Interface Materials from Fujipoly America|SARCON Thermal
distribution effects on the effective thermal conductivity of microgranular
Management Components, 22-Jan-2016. [Online]. Available: <http://
porous media, J. Colloid Interface Sci. 311 (2) (2007) 562–570.
www.fujipoly.com/usa/products/sarcon-thermal-management-components/>.
[31] Y. Agari, A. Ueda, M. Tanaka, S. Nagai, Thermal conductivity of a polymer filled
[Accessed: 22-Jan-2016].
with particles in the wide range from low to super-high volume content, J.
[8] X. Huang, P. Jiang, T. Tanaka, A review of dielectric polymer composites with
Appl. Polym. Sci. 40 (5–6) (1990) 929–941.
high thermal conductivity, Electr. Insul. Mag. IEEE 27 (4) (2011) 8–16.
[32] D. Ganapathy, K. Singh, P.E. Phelan, R. Prasher, An effective unit cell approach
[9] J.C. Maxwell, A Treatise on Electricity and Magnetism, third ed., Dover, New
to compute the thermal conductivity of composites with cylindrical particles,
York, 1954.
J. Heat Transfer 127 (6) (2005) 553–559.
[10] H.T. Davis, L.R. Valencourt, C.E. Johnson, Transport processes in composite
[33] C. Yue, Y. Zhang, Z. Hu, J. Liu, Z. Cheng, Modeling of the effective thermal
media, J. Am. Ceram. Soc. 58 (9–10) (1975) 446–452.
conductivity of composite materials with FEM based on resistor networks
[11] K. Pietrak, T.S. Wiśniewski, A review of models for effective thermal
approach, Microsyst. Technol. 16 (4) (2010) 633–639.
conductivity of composite materials, J. Power Technol. 95 (1) (2014) 14–24.
[34] A. Devpura, P.E. Phelan, R.S. Prasher, Percolation theory applied to the analysis
[12] L. Rayleigh, LVI. On the influence of obstacles arranged in rectangular order
of thermal interface materials in flip-chip technology, in: The Seventh
upon the properties of a medium, Lond. Edinb. Dublin Philos. Mag. J. Sci. 34
Intersociety Conference on Thermal and Thermomechanical Phenomena in
(211) (1892) 481–502.
Electronic Systems, ITHERM 2000, vol. 1, 2000.
[13] H. Levine, The effective conductivity of a regular composite medium, IMA J.
[35] B. Dan, B.G. Sammakia, G. Subbarayan, S. Kanuparthi, S. Mallampati, The study
Appl. Math. 2 (1) (1966) 12–28.
of the polydispersivity effect on the thermal conductivity of particulate
[14] R.C. McPhedran, D.R. McKenzie, The conductivity of lattices of spheres. I. The
thermal interface materials by finite element method, IEEE Trans. Compon.
simple cubic lattice, Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 359 (1978) 45–
Packag. Manuf. Technol. 3 (12) (2013) 2068–2074.
63.
[36] R.T. Bonnecaze, J.F. Brady, A method for determining the effective conductivity
[15] D.R. McKenzie, R.C. McPhedran, G.H. Derrick, The conductivity of lattices of
of dispersions of particles, Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 430
spheres. II. The body centred and face centred cubic lattices, Proc. R. Soc. Lond.
(1990) 285–313.
A: Math. Phys. Eng. Sci. 362 (1978) 211–232.
[37] S. Durmaz, A numerical study on the effective thermal conductivity of
[16] A.S. Sangani, A. Acrivos, The effective conductivity of a periodic array of
composite material (Ph.D. thesis), Graduate School of Natural and Applied
spheres, Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 386 (1983) 263–275.
Sciences of Dokuz Eylul University, 2004.
[17] A.A. Zick, Heat conduction through periodic arrays of spheres, Int. J. Heat Mass
[38] A. Bjorneklett, L. Haukeland, J. Wigren, H. Kristiansen, Effective medium theory
Transfer 26 (3) (1983) 465–469.
and the thermal conductivity of plasma-sprayed ceramic coatings, J. Mater. Sci.
[18] J.B. Keller, Conductivity of a medium containing a dense array of perfectly
29 (15) (1994) 4043–4050.
conducting spheres or cylinders or nonconducting cylinders, J. Appl. Phys. 34
[39] C. Shao, H. Bao, A molecular dynamics investigation of heat transfer across a
(4) (1963) 991–993.
disordered thin film, Int. J. Heat Mass Transfer 85 (2015) 33–40.
[19] D.P.H. Hasselman, L.F. Johnson, Effective thermal conductivity of composites
[40] A. Devpura, Size effects on the thermal conductivity of polymers laden with
with interfacial thermal barrier resistance, J. Compos. Mater. 21 (6) (1987)
highly conductive filler particles, Microscale Thermophys. Eng. 5 (3) (2001)
508–515.
177–189.
[20] A.G. Every, Y. Tzou, D.P.H. Hasselman, R. Raj, The effect of particle size on the
[41] R.C. McPhedran, G.W. Milton, Bounds and exact theories for the transport
thermal conductivity of ZnS/diamond composites, Acta Metall. Mater. 40 (1)
properties of inhomogeneous media, Appl. Phys. A 26 (4) (1981) 207–220.
(1992) 123–129.
[42] S.A. Putnam, D.G. Cahill, B.J. Ash, L.S. Schadler, High-precision thermal
[21] H. Cheng, S. Torquato, Effective conductivity of periodic arrays of spheres with
conductivity measurements as a probe of polymer/nanoparticle interfaces, J.
interfacial resistance, Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 453 (1997)
Appl. Phys. 94 (10) (2003) 6785–6788.
145–161.

You might also like