Sei sulla pagina 1di 16

http://jtc.sagepub.

com/
Composite Materials
Journal of Thermoplastic
http://jtc.sagepub.com/content/19/4/441
The online version of this article can be found at:

DOI: 10.1177/0892705706062203
2006 19: 441 Journal of Thermoplastic Composite Materials
Dilek Kumlutas and Ismail H. Tavman
Filled Polymer Composites
A Numerical and Experimental Study on Thermal Conductivity of Particle


Published by:
http://www.sagepublications.com
at:
can be found Journal of Thermoplastic Composite Materials Additional services and information for





http://jtc.sagepub.com/cgi/alerts Email Alerts:

http://jtc.sagepub.com/subscriptions Subscriptions:
http://www.sagepub.com/journalsReprints.nav Reprints:

http://www.sagepub.com/journalsPermissions.nav Permissions:

http://jtc.sagepub.com/content/19/4/441.refs.html Citations:

at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
A Numerical and Experimental Study
on Thermal Conductivity of Particle
Filled Polymer Composites
DILEK KUMLUTAS AND ISMAIL H. TAVMAN*
Dokuz Eylul University, Mechanical Engineering Department
Bornova, Izmir 35100, Turkey
ABSTRACT: In this study, thermal conductivity of particle filled polymer
composites is investigated numerically and experimentally. In the numerical study,
the finite-element program ANSYS is used to calculate the thermal conductivity of
the composite by using the results of the thermal analysis. Three-dimensional models
are used to simulate the microstructure of composite materials for various filler
concentrations at various ratios of thermal conductivities of filler to matrix material.
The models used to simulate particle filled composite materials are cubes in a cube
lattice array and spheres in a cube lattice array. A modified hot wire method is used
to measure the thermal conductivity of the composites consisting of a high-density
polyethylene (HDPE) matrix filled with tin particles up to 16% by volume. The
experimentally measured thermal conductivities are compared with numerically
calculated ones by using the spheres in cube model and also with the already exist-
ing theoretical and empirical models. At low particle content, up to 10% of volume
content of tin filler, numerical estimation and all other models except for the
Cheng and Vachon model, predict well the thermal conductivity of the composite.
For more heavily filled composites there is an exponential increase in thermal
conductivity and most of the models fail to predict thermal conductivity in this region.
KEY WORDS: polymer composite, thermal conductivity, finite-element analysis,
particle filled.
INTRODUCTION
C
OMMONLY USED POLYMER materials such as polyethylene (PE),
polypropylene (PP), polyvinyl chloride (PVC) or polyamide are good
Journal of THERMOPLASTIC COMPOSITE MATERIALS, Vol. 19July 2006 441
0892-7057/06/04 044115 $10.00/0 DOI: 10.1177/0892705706062203
2006 SAGE Publications
*Author to whom correspondence should be addressed. E-mail: ismail.tavman@deu.edu.tr
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
electrical and thermal insulators. Recent applications of polymers as heat
sinks in electronic packaging require new composites with relatively high
thermal conductivity. Improved thermal conductivity in polymers may be
achieved either by molecular orientation or by the addition of conductive
fillers.
It is well known that thermal transport increases significantly in the
direction of orientation and decreases slightly in the direction perpendicular
to the orientation. Considerable work has been done about this subject
beginning in the 1960s [15]. In a more recent publication by Griesinger,
Hurler, and Pietralla [6] thermal conductivity of low-density PE was seen
to increase from 0.35 for an isotropic sample, to the value of 50 W/mK for
a sample with an orientation ratio of 50. This value of thermal conductivity
is in the range of thermal conductivity for steel.
However, it is not always possible to manufacture polymer objects of a
desired shape with a prescribed orientation value and in the desired dir-
ection. Hence a more practical method to increase the thermal conductivity
of a polymer is needed. The addition of thermally conductive particles or
short fibers to the polymers during the injection molding process seems to be
a good method to obtain conductive polymers. Such filled polymers with
higher thermal conductivities than unfilled ones are becoming a more
important area of study because of the wide range of applications,
e.g., in electronic packaging in applications with decreasing geometric
dimensions and increasing output of power, like in computer chips or in
electronic packaging [79].
There are many experimental as well as numerical and analytical model
studies on thermal conductivity of filled polymer composites [1022]. The
fillers most frequently used are aluminum particles, copper particles, brass
particles, short carbon fiber, carbon particles, graphite, aluminum nitrides,
and magnetite particles. In an experimental study on aluminum-filled high-
density polyethylene (HDPE) composites by Tavman [20], it has been
observed that thermal conductivity increased from 0.543 W/mK for pure
HDPE samples to the value of 3.671 W/mK for composite samples filled
with 33% by volume aluminum particles. In a more recent study by Sofian
et al. [21], thermal properties such as thermal conductivity, thermal
diffusivity, and specific heat of metal (copper, zinc, iron, and bronze)
powder filled HDPE composites are investigated experimentally in the range
of filler content 024% by volume. They observed a moderate increase in
thermal conductivity up to 16% of metal powder filler content. In this
region most of the predictive models for thermal conductivity of two-phase
systems are applicable. At higher filler contents, the filler particles tend to
form agglomerates and conductive chains resulting in a rapid increase in
thermal conductivity. In a more recent study by Tekce et al. [22] thermal
442 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
conductivity of lamellar and spherical copper powder filled polyamide-6
(PA 6) composites are investigated. From this study the strong influence
of the shape factor of fillers on thermal conductivity of the composite
may be noticed: for filler volume fraction of 50% thermal conductivity
has increased from 0.32 W/mK for pure PA 6 to 7.29 and 2.09 W/mK
for lamellar and spherical copper powders filled composites respectively.
A numerical approach to predict the effective thermal conductivity of gran-
ular or fibrous reinforced composite materials was proposed by Veyret et al.
[17]. In their study, using a finite-element formulation, calculation was
carried out on two- and three-dimensional geometric spaces. The results
obtained from this calculation were compared to theoretical results found
in prior literature.
Published values of thermal conductivities of the same filler materials
in the same polymer matrices vary more or less for each study; this is mainly
due to the mode of sample preparation as some samples are prepared
by compression molding, some others by extrusion and injection molding.
The size and shape of the filler and also the interconnectivity of the filler
particles in the matrix may be other factors which influence the thermal
conductivities of the composites. In a recent research Weidenfeller et al. [23]
studied the effect of the interconnectivity of the filler particles and its
important role in the thermal conductivity of the composites. They prepared
PP samples with different commercially available fillers by extrusion
and injection molding using various volume fractions of filler content to
systematically vary density and thermal transport properties of these
composites. Surprisingly, they measured that the thermal conductivity of the
PP has increased from 0.27 up to 2.5 W/mK with 30 vol% talc in the
PP matrix, while the same matrix material containing the same volume
fraction of copper particles had a thermal conductivity of only 1.25 W/mK,
despite the fact that copper particles have a thermal conductivity
approximately 40 times greater than that of talc particles. They related
these results to a complete interconnectivity achieved for talc fillers in PP,
while copper particles in PP show a very poor interconnectivity.
In this study, the effective thermal conductivity of particle and fiber
reinforced composites is investigated numerically by using the finite-
element program ANSYS. Two- and three-dimensional models are used
to simulate the microstructure of composite materials for various filler
concentrations at various ratios of thermal conductivities of filler to matrix
(k
f
/k
m
). Furthermore, the effective thermal conductivity of tin particles
filled HDPE composites is calculated numerically as a function of filler
concentration. The results obtained from this calculation are compared with
the existing theoretical and empirical models, and also with experimental
results.
Thermal Conductivity of Particle Filled Polymer Composites 443
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
THERMAL CONDUCTIVITY MODELS
In this section, several predictive models for thermal conductivity of
composites are listed with a brief description of their basic assumptions.
Many theoretical and empirical models have been proposed to predict the
effective thermal conductivity of two-phase mixtures. Comprehensive review
articles have discussed the applicability of many of these models [24,25].
For a two-component composite, the simplest alternatives would be with
the materials arranged in either parallel or series with respect to heat flow,
which gives the upper or lower bounds of effective thermal conductivity.
For the parallel conduction model:
k
c
1 k
m
k
f
1
where, k
c
, k
m
, k
f
are the thermal conductivities of the composite, the matrix
and the filler respectively and is the volume fraction of filler.
For series conduction model:
1
k
c

1
k
m

k
f
2
In the case of the geometric mean model, the effective thermal
conductivity of the composite is given by:
k
c
k

f
:k
1
m
3
Tsao [26] derived an equation relating the two-phase solid mixture thermal
conductivity to the conductivity of the individual components and to two
parameters which describe the spatial distribution of the two phases.
By assuming a parabolic distribution of the discontinuous phase in the
continuous phase, Cheng and Vachon [27] obtained a solution to Tsaos [26]
model that did not require knowledge of additional parameters. The constants
of this parabolic distribution were determined by analysis and presented as
a function of the discontinuous-phase volume fraction. Thus, the equivalent
thermal conductivity of the two-phase solid mixture was derived in terms
of the distribution function, and the thermal conductivity of the constituents.
For, k
f
>k
m
1
k
c

Ck
f
k
m
k
m
Bk
f
k
m

p
ln

k
m
Bk
f
k
m

p
B=2

Ck
f
k
m

k
m
Bk
f
k
m

p
B=2

Ck
f
k
m

p
1 B
k
m
4
444 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
where,
B

3=2
p
C 4

2=3
p
Lewis and Nielsen [28] derived a semi-theoretical model by a modifica-
tion of the HalpinTsai equation [29] to include the effect of the shape of the
particles and the orientation or type of packing for a two-phase system:
k
c
k
m
1 AB
1 B

5
where,
B
k
f
=k
m
1
k
f
=k
m
A
1
1
m

2
m

The values of A and


m
for many geometric shapes and orientation are given
in Tables 1 and 2.
Using potential theory, Maxwell [30] obtained an exact solution for the
conductivity of randomly distributed and non-interacting homogeneous
spheres in a homogeneous medium:
k
c
k
m
k
f
2k
m
2k
f
k
m

k
f
2k
m
k
f
k
m


6
This model predicts fairly well the effective thermal conductivities at low
filler concentrations; whereas for high filler concentrations, particles begin
Table 1. Value of A for various systems.
Type of dispersed phase Direction of heat flow A
Cubes Any 2.0
Spheres Any 1.50
Aggregates of spheres Any (2.5/
n
) 1
Randomly oriented rods aspect ratio2 Any 1.58
Randomly oriented rods aspect ratio4 Any 2.08
Randomly oriented rods aspect ratio6 Any 2.8
Randomly oriented rods aspect ratio10 Any 4.93
Randomly oriented rods aspect ratio15 Any 8.38
Uniaxially oriented fibers Parallel to fibers 2L/D
Uniaxially oriented fibers Perpendicular to fibers 0.5
Thermal Conductivity of Particle Filled Polymer Composites 445
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
to touch each other and form conductive chains in the direction of heat
flow, so that this model underestimated the value of effective thermal
conductivities in this region.
Agari and Uno [31] propose a new model for filled polymers, which takes
into account parallel and series conduction mechanisms. According to this
model, the expression that governs the thermal conductivity of the
composite is:
log k
C
C
2
log k
f
1 logC
1
k
m
7
where, C
1
, C
2
are experimentally determined constants of order unity. C
1
is
a measure of the effect of the particles on the secondary structure of the
polymer, like crystallinity and the crystal size of the polymer. C
2
measures
the ease of the particles to form conductive chains. The more easily particles
are gathered to form conductive chains, the more thermal conductivity of
the particles contributes to change in thermal conductivity of the composite
and C
2
becomes closer to 1. Later, they modified the model to take into
account the shape of the particles [32]. Generally, this semiempirical model
seems to fit the experimental data well. However, adequate experimental
data is needed for each type of composite in order to determine the necessary
constants.
EXPERIMENTAL
Sample Preparation
Samples are prepared by the mold compression process. The
matrix material is a HDPE in powder form, with a solid density of
0.968 g/cm
3
and a melt index of 5.8 g/10 min. Its thermal conductivity
Table 2. Value of /
m
for various systems.
Shape of particle Type of packing /
m
Spheres Hexagonal close 0.7405
Spheres Face centered cubic 0.7405
Spheres Body centered cubic 0.60
Spheres Simple cubic 0.524
Spheres Random close 0.637
Rods or fibers Uniaxial hexagonal close 0.907
Rods or fibers Uniaxial simple cubic 0.785
Rods or fibers Uniaxial random 0.82
Rods or fibers Three-dimensional random 0.52
446 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
at 38

C is 0.554 W/mK. The metallic filler tin is in the form of fine powder,
with particles approximately spherical in shape and particle size in the range
of 2040 m. The solid density of tin is 7.28 g/cm
3
and its thermal conductivity
64 W/mK. HDPE and tin powders are mixed at various volumetric
concentrations. The mixed powder is then melted under pressure in a
mold and solidified in the mold by air-cooling. The process conditions are
molding temperature of 185

C, pressure of 4 MPa. The resulting samples


for thermal conductivity measurements are rectangular in shape of 100 mm
length, 50 mm width, and 17 mm thickness.
Homogeneity of the samples is examined using a light microscope
(Figure 1(a) and (b)). Microscopic study shows that the tin particles are
nearly spherical in shape at low particle content, <10. Tin particles are
found to be uniformly distributed in HDPE matrix with no voids in the
structure. At higher filler contents, as aggregates of particles are formed,
the shape of the particles cannot be taken as spherical.
Measurement Technique
Thermal conductivity measurements of pure HDPE as well as tin-
filled HDPE composites are carried out with the Shotherm QTM thermal
conductivity meter using a modified hot wire technique. A thin straight
wire through which a constant electric current is passed, generating constant
heat (Q) per unit length of wire, per unit time (t), is placed between
two rectangular shaped materials, the first one is an insulating material of
known thermal conductivity which is a part of the measuring probe and
the second one is the sample for which the thermal conductivity has to be
(a) (b)
60 m
60 m
Figure 1. Microscopic photographs of HDPE filled with tin: (a) 10 volume percent and (b) 16
volume percent.
Thermal Conductivity of Particle Filled Polymer Composites 447
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
measured (Figure 2). A constant power is supplied to the heater element
and the temperature rise (T) of the heating wire is measured by a
thermocouple and recorded with respect to time during a short heating
interval, the thermal conductivity (k) of the sample is calculated from
the temperaturetime (Tt) record and power input (Q) according to the
equation:
k F
Qlnt
2
=t
1

T
2
T
1
H 8
where, F and H are specific constants of the probe to be determined
with materials of known thermal conductivities. By this method, the thermal
conductivity is measured with an accuracy of 5% and reproducibility of
2%. For each specimen the thermal conductivity is measured five times
in a temperature range from about 5

C to 70

C and the mean values


are reported.
NUMERICAL ANALYSIS
Using the finite-element program ANSYS, thermal analysis was carried
out for the conductive heat transfer. In order to make a thermal analysis,
three-dimensional models have been used to simulate the microstructure
of composite materials for various filler concentrations at various k
f
/k
m
.
They consist of cubes in a cube lattice array and spheres in a cube lattice
array, (Figure 3(a) and (b)). Furthermore, the effective thermal conductivity
of HDPE filled with tin particles up to 16% by volume is numerically
calculated by using the sphere in cube model.
In the numerical analysis of the heat conduction problem, the temp-
eratures at the nodes along the surfaces ABCD and EFGH (Figure 4)
are prescribed as T
1
and T
2
. The other surfaces parallel to the direction
of the heat flow are all assumed adiabatic. The temperatures at the nodes
Heating wire
Composite
sample
Material of
known k
Thermocouple
Figure 2. Thermal conductivity measuring probe.
448 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
in the interior region and on the adiabatic boundaries are unknown.
These temperatures are obtained with the finite-element program package
ANSYS. Effective thermal conductivity of the composites is calculated by
using the results of the thermal analysis.
For an elementary three-dimensional cell with the dimensions of L
x
(along
the x-axis), L
y
(along the y-axis), and L
z
(along the z-axis), the thermal
conductivity is calculated using the following relation:
k
c
T
cell
L
z

1
S
X
i
X
j
k
ij
x
i
y
j
@T
ij
@z

9
L
z
Q
A B
C
D
T
2
T
1
L
x
L
y
k
m
x
y
z
k
f
E
H G
F
s
ij
Figure 4. Boundary conditions.
(a) (b)
Figure 3. Three-dimensional finite-element models for 44% particle concentration: (a)
sphere in cube and (b) cube in cube.
Thermal Conductivity of Particle Filled Polymer Composites 449
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
with
s
ij
x
i
y
j
,
X
i
X
j
s
ij
S and k
ij

k
m
in the continuous phase
k
f
in the inclusions

RESULTS AND DISCUSSION
The effective thermal conductivities are determined numerically for
different phase conductivity ratios (110 000), for different phase quantities
(4, 27, 44% by volume for the three-dimensional models). Figure 5 shows
the effective thermal conductivity variations for spheres in cube and cubes in
cube systems for different conductivity ratio k
f
/k
m
and concentration.
Similar behavior can be noted for both types of model. With large
conductivity ratio k
f
/k
m
values, a horizontal asymptote is observed. As seen
from this figure, for low concentrations of filler content, the effective
thermal conductivity values calculated are nearly the same for both of
the three-dimensional models, whereas these values depend on the geometry
of the filler for large concentrations, sphere in cube model predicting
bigger values. This may be explained by the fact that spherical particles
begin to touch each other at lower concentrations compared with the cubic
particles.
Thermal conductivity measurements are performed on HDPE filled with
tin particles up to a volumetric fraction of 16%, in a temperature range
0
1
2
3
4
5
1 10 100 1000 10000
Sphere in cube (4%)
Sphere in cube (27%)
Sphere in cube (44%)
Cube in cube (4%)
Cube in cube (27%)
Cube in cube (44%)
k
f
/k
m
k
c
/
k
m
Figure 5. Variation of the effective thermal conductivity according to the ratio k
f
/k
m
for
various particle concentrations and geometries.
450 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
from about 5

C to 70

C. Measured thermal conductivities for various


particle concentrations are plotted as a function of temperature in Figure 6.
In accordance with literature values, thermal conductivity decreases with
increasing temperature, as the measurement range is well above the glass
transition temperature, which is 110

C for HDPE.
Furthermore, thermal conductivities of HDPE filled with tin particles
up to 16% by volume are numerically estimated by using the spheres in
cube model and the numerical results are compared with the experimental
results and also with existing theoretical and empirical models in Figure 7.
It is seen that the numerical results, experimental values and all the models
except for the Cheng and Vachon model are close to each other at low
particle content, <10, as the particles are dispersed in the polymer matrix
and they are not interacting with each other. For particle content greater
than 10% by volume, conductive chains are formed by tin particles causing
a large increase in effective thermal conductivity of the composite. Agari
and Uno [31] and Cheng and Vachon [27] models predictions of thermal
conductivities are more accurate in this region. Especially, the Agari and
Uno [31] model predicts quite well the thermal conductivities of the
composite in the whole range of tin particle content. In the Agari and Uno
[31] model, values of the coefficients C
1
and C
2
are calculated by using the
linear regression equation of the plot of the logarithm of the thermal
0.4
0.6
0.8
1
1.2
1.4
1.6
10 0 10 20 30 40 50 60 70
Temperature (C)
T
h
e
r
m
a
l

c
o
n
d
u
c
t
i
v
i
t
y

(
W
/
m
K
)
Pure HDPE 3% Sn 8% Sn
10% Sn 13% Sn 16% Sn
Figure 6. Measured thermal conductivity of HDPE filled with tin particles as a function of
temperature.
Thermal Conductivity of Particle Filled Polymer Composites 451
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
conductivities of the experimental data against volume contents of the
particles. In the present study, C
1
0.936 and C
2
0.897. As it uses the
actual experimental data to fit a curve, it is natural that this model predicts
best the effective thermal conductivities of filled systems in the whole range.
For particle content greater than 10% by volume, the Lewis and Nielsen [28]
model with A3 and
m
0.637 agrees fairly well with experimental results
0.5
0.6
0.7
0.8
0.9
1
1.1
1.2
1.3
1.4
0 5 10 15 20
Sn%
E
f
f
e
c
t
i
v
e

t
h
e
r
m
a
l

c
o
n
d
u
c
t
i
v
i
t
y

(
W
/
m
K
)
Numerical
Experimental
Cheng and Vachon
Lewis and Nielsen (A=3,
m
=0.637)
Agari and Uno
Lewis and Nielsen (A=1.5,
m
=0.637)
Maxwell
Figure 7. Comparison of the numerical results, models, and experimental thermal
conductivity values of tin-filled HDPE for various tin particle concentration at 38

C.
452 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
as aggregates of particles are formed and they form thermal bridges in the
direction of heat flow. The numerical results obtained from this study agree
very well the Maxwell model [30], as the assumptions of spherical parti-
cles uniformly distributed and not interacting with each other is the same
for both cases.
CONCLUSIONS
The addition of conductive fillers such as metallic and carbon particle or
short fiber fillers in the polymer matrix is an effective way to increase
thermal conductivity of the polymers, as required by several industrial
applications. All the theoretical, empirical as well as numerical models fail
to predict thermal conductivity of filled polymer composites in the whole
range of filler content. However, up to 10% of filler content, the numerical
model developed in this study and some models like Maxwells model and
the Lewis and Nielsen model predict fairly well the thermal conductivity of
the polymer composite. The Agari and Uno model uses the experimental
results to derive two constants of the model, so it is natural that it fits the
experimental curve. To make adequate use of the filled polymer composites
especially in heat sink applications, experimental results are necessary for
the specific type of composite in question. In this study, for tin particle filled
HDPE, thermal conductivity increases from 0.554 W/mK for pure HDPE
samples to 0.681 W/mK and 1.116 W/mK for 8% and 16% filler content by
volume respectively, which represents 23% and 101% increase. Thus, a
rapid exponential increase in thermal conductivity for filler contents more
than 10% by volume is observed.
REFERENCES
1. Hennig, J. and Knappe, W. (1964). Anisotropy of Thermal Conductivity in Stretched
Amorphous Linear Polymers and in Strained Elastomers, Journal of Polymer Science
Part C, 6: 167174.
2. Hansen, D. and Ho, C. (1965). Thermal Conductivity of High Polymers, Journal of Polymer
Science Part A, 3(2): 659670.
3. Peng, S. and Landel, R. (1975). Induced Anisotropy of Thermal Conductivity of Polymer
Solids under Large Strains, Journal of Applied Polymer Science, 19(1): 4968.
4. Choy, C.L. and Young, K. (1977). Thermal Conductivity of Semicrystalline PolymersA
Model, Polymer, 18(8): 769776.
5. Tavman, I. (1991). Thermal Anisotropy of Polymers as a Function of their Molecular
Orientation, Experimental Heat Transfer, Fluid Mechanics, and Thermodynamics,
pp. 15621568, Elsevier.
6. Griesinger, A., Hurler, W. and Pietralla, M. (1997). A Photothermal Method with Step
Heating for Measuring the Thermal Diffusivity of Anisotropic Solids, International Journal
of Heat and Mass Transfer, 40(13): 30493058.
Thermal Conductivity of Particle Filled Polymer Composites 453
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
7. Wong, C.P. and Bollampally, R.S. (1999). Thermally Conductivity, Elastic Modulus,
and Coefficient of Thermal Expansion of Polymer Composites Filled with Ceramic
Particles for Electronic Packaging, Journal of Applied Polymer Science, 74(14): 33963403.
8. Lu, X. and Xu, G. (1997). Thermally Conductive Polymer Composites for Electronic
Packaging, Journal of Applied Polymer Science, 65(13): 27332738.
9. Xu, Y., Chung, D.D.L. and Mroz, C. (2001). Thermally Conducting Aluminum Nitride
Polymer-matrix Composites, Composites: Part A, 32(12): 17491757.
10. Mamunya, Y.P., Davydenko, V.V., Pissis, P. and Lebedev, E.V. (2002). Electrical
and Thermal Conductivity of Polymers Filled with Metal Powders, European Polymer
Journal, 38(9): 18871897.
11. Saxena, N.S., Pradeep, P., Mathew, G., Thomas, S., Gustafsson, M. and
Gustafsson, S.E. (1999). Thermal Conductivity of Styrene Butadiene Rubber
Compounds with Natural Rubber Prophylactics Waste as Filler, European Polymer
Journal, 35(9): 16871693.
12. Ishida, H. and Rimdusit, S. (1998). Very High Thermal Conductivity Obtained by
Boron Nitride-filled Polybenzoxazine, Thermochimica Acta, 32(12): 177186.
13. Tavman, I.T. (1997). Thermal and Mechanical Properties of Copper Powder Filled Poly
(ethylene) Composites, Powder Technology, 91(1): 6367.
14. Bjorneklett, A., Halbo, L. and Kristiansen, H. (1992). Thermal Conductivity of Epoxy
Adhesives Filled with Silver Particles, International Journal of Adhesion and Adhesives,
12(2): 99104.
15. Agari, Y., Ueda, A., Tanaka, M. and Nagai, S. (1990). Thermal Conductivity of a
Polymer Filled with Particles in the Wide Range from Low to Super-high Volume Content,
Journal of Applied Polymer Science, 40(56): 929941.
16. Kumlutas, D., Tavman, I.H. and Coban, M.T. (2003). Thermal Conductivity of
Particle Filled Polyethylene Composite Materials, Composites Science and Technology,
63(1): 113117.
17. Veyret, D., Cioulachtjian, S., Tadrist, L. and Pantaloni, J. (1993). Effective Thermal
Conductivity of a Composite Material: A Numerical Approach, Transactions of the
ASME- Journal of Heat Transfer, 115: 866871.
18. Mottram, J.T. and Taylor, R. (1987). Thermal Conductivity of Fibre/Phenolic Resin
Composites. Part II: Numerical Evaluation, Composites Science and Technology, 29(3):
211232.
19. Onyejekwe, O.O. (2002). Heat Conduction in Composite Media: A Boundary Integral
Approach, Computers & Chemical Engineering, 26(11): 16211632.
20. Tavman, I.H. (1996). Thermal and Mechanical Properties of Aluminum Powder-
filled High-density Polyethylene Composites, Journal of Applied Polymer Science, 62(12):
21612167.
21. Sofian, N.M., Rusu, M., Neagu, R. and Neagu, E. (2001). Metal Powder-filled
Polyethylene Composites. V. Thermal Properties, Journal of Thermoplastic Composite
Materials, 14(1): 2033.
22. Tekce, H.S., Kumlutas, D. and Tavman, I.H. (2004). Determination of the Thermal
Properties of Polyamide-6 (Nylon-6)/Copper Composite by Hot Disk Method,
In: Proceedings of the 10th Denizli Material Symposium, Denizli, April 2004, pp. 296304.
23. Weidenfeller, B., Ho fer, M. and Schilling, F.R. (2004). Thermal Conductivity, Thermal
Diffusivity, and Specific Heat Capacity of Particle Filled Polypropylene, Composites Part
A: Applied Science and Manufacturing, 35(4): 423429.
24. Ott, H.J. (1981). Thermal Conductivity of Composite Materials, Plastic and Rubber
Processing and Application, 1(1): 924.
25. Progelhof, R.C., Throne, J.L. and Ruetsch, R.R. (1976). Methods of Predicting the
Thermal Conductivity of Composite Systems, Polymer Engineering and Science, 16(9):
615625.
454 D. KUMLUTAS AND I. H. TAVMAN
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from
26. Tsao, T.N.G. (1961). Thermal Conductivity of Two Phase Materials, Industrial and
Engineering Chemistry, 53(5): 395397.
27. Cheng, S.C. and Vachon, R.I. (1969). The Prediction of the Thermal Conductivity of
Two and Three Phase Solid Heterogeneous Mixtures, International Journal of Heat Mass
Transfer, 12(3): 249264.
28. Lewis, T. and Nielsen, L. (1970). Dynamic Mechanical Properties of Particulate-filled
Polymers, Journal of Applied Polymer Science, 14(6): 14491471.
29. Halpin, J.C. (1969). Stiffness and Expansion Estimates for Oriented Short Fiber
Composites, Journal of Composite Materials, 3(1): 732734.
30. Maxwell, J.C. (1954). A Treatise on Electricity and Magnetism, 3rd edn, Ch.9. Dover Inc.,
New York, NY.
31. Agari, Y. and Uno, T. (1986). Estimation on Thermal Conductivities of Filled Polymers,
Journal of Applied Polymer Science, 32(7): 57055712.
32. Agari, Y., Ueda, A. and Nagai, S. (1991). Thermal Conductivity of Polyethylene Filled
with Disoriented Short-cut Carbon Fibers, Journal of Applied Polymer Composite Science,
43(6): 11171124.
Thermal Conductivity of Particle Filled Polymer Composites 455
at UNIV ESTADUAL CAMPINAS BIBLIO on May 18, 2011 jtc.sagepub.com Downloaded from

Potrebbero piacerti anche