Você está na página 1de 10

Composires Engmeenng, Vol. 5, No. 9. pp. I177-1 IR6.

1995
Copyright fs: 1995 Elsewer Science Ltd
PrInted in Great Britain. All rights reserved
0961.9526/95 $9.50,
IH)

Pergamon

O%l-9526(94) 00106-5

THERMAL
MADE

CONDUCTIVITY
OF COMPOSITE MATERIALS
FROM PLAIN WEAVES AND 3-D WEAVES
Yasser Gowayed and Jhy-Cherng Hwang

Department of Textile Engineering, Auburn University, Auburn, AL 36849, U.S.A.


(Received 8 September 1994;final version accepted 16 November

1994)

Abstract-In
this research activity, thermal conductivity of textile composites made from plain
weaves and 3-D XYZ weaves is quantified. Plain weave composites are made from E-glass,
KevlaP and AS4 graphite fibers and epoxy resin. 3-D woven composites are manufactured from
Toho graphite fibers and epoxy resin. The effect of fiber type and fiber volume fraction on the
thermal conductivity of textile composites is investigated.
The fabric Geometry Model is adapted to calculate the thermal conductivity of textile
composites. Results using this tool are compared with experimental data and predictions using the
Graphical Integrated Numerical Analysis. The analytical approach proved to be a good engineering tool to predict the thermal conductivity of textile composites.

INTRODUCTION

Published literature is rich with investigations of mechanical properties of composites.


Fewer accounts are oriented towards thermal properties. Currently, the area of thermal
conductivity is getting more attention to expand the applications of composites. Most of
the research work in this area is directed towards developing prediction models and
experimental work for thermal conductivity of unidirectional and particulate composites,
see for example, Rayleigh (1894), Maxwell (1904), Springer and Tsai (1967), Behrens
(1968), Donea (1972), Han and Cosner (1981), Brennan et al. (1982), Tzadka and
Schulgasser (1983), Tawil et al. (1985), James el a/. (1987) Hasselman et al. (1987) and
Havis et al. (1989).
Research oriented towards experimentally quantifying thermal conductivity of
various textile composite structures is very limited (Lubin, 1981; Stuart, 1990). Analytical
prediction models are also limited and have a high degree of complexity. Dasgupta and
Agrawal (1992) introduced an asymptotic scheme to predict the orthotropic thermal
conductivity of plain weave composites employing a 3-D thermal resistance network.
Also, Gowayed et al. (1995) presented a Graphical Integrated Numerical Analysis Model
(GINA) based on hybrid finite element model coupled with a geometrical representation
of the fabric preform to predict the thermal conductivity of textile composites with
arbitrary preform structures.
In this paper, an experimental investigation of the thermal conductivity of textile
composites made from plain weaves and 3-D XYZ weaves is carried out. The effect of
fiber type and fiber volume fraction on the thermal conductivity of textile composites is
investigated. Moreover, an easy to apply averaging technique (Kregers and Melbardis,
1978; Whyte, 1986; Pastore and Gowayed, 1994) is adapted to calculate the thermal
conductivity of textile composites. Results using this tool are compared with experimental
data and predictions obtained using the Graphical Integrated Numerical Analysis
(Gowayed, 1992; Gowayed et al., 1995).
EXPERIMENTAL

WORK

Test apparatus
The thermal conductivity chamber used in this study is a k-Matic 75 thermal conductivity instrument manufactured by Holometrix, Inc. The design of this instrument is in
accordance with ASTM Test Method C518 (1985) for steady state heat flux measurements
and thermal transmission properties by means of a heat flow meter.
1177

Y. Gowayed and J.-C. Hwang

1178

Guiderod

Movable base

plate

Fig. 1. Schematic cross-sectional view of the thermal conductivity testing chamber.

Figure 1 illustrates a cross-section of the testing chamber. The testing chamber is


constructed of two parts: a stationary upper section and a movable lower section to
accommodate specimens with various thicknesses. The upper section consists of a
refrigerated heat sink (HS) separated by insulation (INSl) from a controlled heater plate
(CHl) and an aluminum surface plate (SPl). The lower section consists of a heat flow
meter (HF) (thermal transducer), with a 10 x 10 cm contact area, to measure the heat flux
separated from the lower surface plate (SP2) by a thin layer of thermal insulation (INS2).
This is followed by a heater plate (CH2). A thermal-couple (TCl) and (TC2) is installed
at both sides of the specimen to measure the heat difference AT.
The refrigerated heat sink is a thermally activated safety switch which serves to interrupt the power supply to the heater circuits if, for any reason, the heat sink temperature
(TC3) rises above 2C during a test.
On the right side of the test chamber is a linear-motion potentiometer for determining
the sample thickness Ax during a test. The potentiometer housing is attached rigidly to
the movable lower section. The test chamber is surrounded by insulation to reduce heat
transfer to or from the elements in the stack.
Test procedure

The thermal conductivity measurement for different composite samples is carried out
by placing the sample between the two surface plates of known temperature. Because of
the temperature difference, heat flows through the sample from the hot side to the cold
side. The quantity of heat flow through the sample is measured by the thermal transducer.
Due to the large size of the heat transducer, the specimen dimensions should be at least
20 x 20 cm to minimize heat loss. The size of the transducer ensures the properties of the
composites over the area under consideration. The large size of the sample allows only the
measurement of heat conduction through the thickness of the sample. In-plane thermal
conductivities are not measured.
The instrument is first calibrated using a specimen with known thermal conductivity.
The thermal conductivity k is measured according to the formula:

(1)
where q = heat flow through the sample measured by the thermal transducer (HF),
A = area of the sample, AT = temperature difference (TC2 - TCl) across sample
thickness Ax.

Thermal conductivity of plain and 3-D weaves

Fig. 2. Micrographic image of balanced plain weave E-glass/epoxy composite with fiber volume
fraction of 34.6%.

Fig. 3. Micrographic image of AScl/epoxy composite with fiber volume fraction of 26.8%.

1179

1180

Y. Gowayed and J.-C. Hwang

Fig. 4. Micrographic image of Kevlar-49/epoxy composite with fiber volume fraction of 42.7%.

Fig. 5. Micrographic image of 3-D XYZ graphite/epoxy


of 29.65%.

composite with fiber volume fraction

Thermal conductivity of plain and 3-D weaves

1181

Materials
Plain weaves. The characteristics of the balanced plain weave fabrics utilized in this
study are as follows:
??

??

??

2 ends/cm 12k E-glass yarns manufactured by PPG Industries. The E-glass fiber
thermal conductivity in the longitudinal and transverse directions, as calculated
from unidirectional test results, is 0.50 (W/mC). Composites tested have total
fiber volume fractions of: 34.6%, 42.6%, 47.0%, 49.7% and 55.7%.
5 ends/cm 3k AS4 graphite yarns manufactured by BP Chemicals. The thermal
conductivities of the fiber (Dasgupta and Agrawal, 1992) are 8.4 (W/mC) along
the fiber direction and 0.84(W/mC) in the transverse direction. Composites
tested have total fiber volume fractions of: 26.8V0, 37.5%, 47.1%, 54.0% and
55.7%.
12 ends/cm lk Kevlar-49@ fibers manufactured by DuPont. The thermal conductivities of the fiber (Kawabata, 1988) are 3.34 (W/mC) along the fiber direction
and 0.212 (W/mC) in the transverse direction. Composites tested have total
fiber volume fractions of: 42.7%, 49.7%, 60.7070, 63.5% and 64.5%.

3-D weaves. Three 3-D XYZ woven fabric preforms made


yarns are tested. The thermal conductivities of the fiber
6.69 (W/m%) along the fiber direction and 0.53 (W/mC) in
Composites tested have total volume fractions of 28.77% (V,,
and V,, = 0.33%), 29.69% (I$, = 13.44%, I$,, = 11.95% and
(5, = 15.6%, V& = 14.0% and vf, = 8.4%).

from 12k Toho graphite


(Kawabata, 1988) are
the transverse direction.
= 15.0070, L$,,= 13.44%
V,, = 4.3%) and 38.0%

Epoxy resin. Ciba-Geigy Araldite epoxy resin and HY 956 hardener are used as the
matrix. The mixing is carried out at room temperature with a ratio of 4 : 1 resin : hardener.
The measured thermal conductivity for pure matrix is 0.196 (W/mC).
Composite manufacturing. Fabrics were manually placed in a 38 x 30 cm mold and
impregnated with epoxy resin. The composite plates were cured in a compression molding
machine at 80C for two hours and then post cured at 120C for one and a half hours.
Composite plates were then cut to 25 x 25 cm for thermal conductivity testing.
Micrographical images of balanced plain weaves E-glass/epoxy, AS4/epoxy, Kevlar49/epoxy composites are shown in Figs 2, 3 and 4, respectively. A micrographic image of
a 3-D XYZ graphite/epoxy composite is shown in Fig. 5.
THEORETICAL ANALYSIS

The Fabric Geometry Model , or the Stiffness Averaging Technique, was originally
developed (Kregers and Melbardis, 1978; Whyte, 1986; Pastore and Gowayed, 1994) to
predict the elastic properties of textile composites. The basic idea behind the Fabric
Geometry Model is to treat the fibers and matrix as a set of composite rods having various
spatial orientations. Each composite rod represents a reinforcing system. The local stiffness
tensor for each of these rods is calculated and rotated in space to fit the global composite
axes. The global stiffness tensors of all the composite rods (i.e. reinforcing systems) are
then superimposed with respect to their relative volume fraction to form the composite
stiffness tensor.
In the current research activity, this approach is adapted to predict the thermal
conductivity of textile composites. The analysis steps are illustrated in Fig. 6.
The first step in the analysis is to calculate the local thermal conductivity tensor for
each reinforcing system. The local conductivity tensor could be represented as follows:

K local
=

k 11

k,,

k,,

k21

k 22

k23

ka

ku I

[ k 31

where k, = symmetric conductivity tensor in the i and j directions (i, j = 1,2, 3).

(2)

Y. Gowayed and J.-C. Hwang

1182

input Material Properties

Thermal conductivifies and Volume fractio

For Each Reinforcing System

Direction cosines

Calculate Transfromation Matrix T


JI
Calculate Global Thermal Conductivity
For Each Reinforcing System
K global
JI
Calculate Thermal Conductivity of Composite
K composite

Fig. 6. Analysis steps for conductivity predictions using a Modified Fabric Geometry Model.

A large number of unidirectional micro-level thermal analysis models, used to calculate Kj,j, do exist in archived literature (Bruggeman, 1935; Springer and Tsai, 1967;
Behrens, 1968; Hashin, 1979). The authors utilized the model developed by Behrens (1968)
due to the large experimental verification data available in literature to support this model.
The next step is to transform the local thermal conductivity tensor Klocalfor each
reinforcing system to fit the directions of the composite global axes. This is done utilizing
the transformation relation:
K global

T- Kloca,T

(3)

where Kgtobat = global thermal conductivity tensor for the composite rod, T = transformation tensor of direction cosines. T could be represented as:

T =

1;

1;

13

rnt

rni

rn:

n:

n: I

[ nT

(4)

The formulation of the transformation matrix depends on the direction cosines li, mi
and n, (i = 1,2, 3). These direction cosines may be observed as components of unit basis
vectors associated with the fiber axis. A pseudo-code is presented in Pastore and Gowayed
(1994) to calculate Ii, mi and Q.
Finally, the composite thermal conductivity matrix K,somposite
is calculated by superimposing the global conductivity tensor for each reinforcing system Kgloba,with respect to
its relative volume fraction as follows:
K composite
= !

i=O

Kglobal,iei

(5)

= relative volume fraction


Kglobal,i = global conductivity of reinforcing system i, 19~
of reinforcing system i and N = total number of reinforcing systems.

where

RESULTS AND DISCUSSION

Plain weaves
Tables l-3 and Figs 7-9 show out-of-plane thermal conductivity test results and
standard deviations, predictions using the modified FGM, detailed in the previous section,

Thermal conductivity of plain and 3-D weaves

1183

Table 1. Experimental results and analytical predictions for out-of-plane


thermal conductivity in (W/mC) for plain weave E-glass/epoxy composites
Volume fraction
(Q)

Prediction

Experiment

34.6
42.6
47.0
49.7
55.7

0.269
0.299
0.313
0.337
0.378

f
+
f
+
f

Mod. FGM

GINA

0.268
0.289
0.30
0.307
0.327

0.214
0.295
0.308
0.316
0.336

0.002
0.005
0.005
0.009
0.032

Table 2. Experimental results and analytical predictions for out-of-plane


thermal conductivity in (W/m(Z) for plain weave AS4/epoxy composites
Prediction
Experiment

Volum(+fjaction
0

26.8
37.5
47.1
54.0
55.7

0.288
0.336
0.373
0.441
0.453

+
f
+
+
+

Mod. FGM

GINA

0.291
0.338
0.387
0.426
0.437

0.304
0.353
0.41 I
0.456
0.469

0.027
0.028
0.028
0.01
0.005

Table 3. Experimental results and analytical predictions for out-of-plane


thermal conductivity in (W/m(Z) for plain weave Kevlar-49/epoxy composites
Prediction
Volume fraction
(%)

Experiment

42.7
49.7
60.7
63.5
64.5

0.236
0.246
0.260
0.268
0.272

-Mod.
*-_*

0.45

+
+
f
f
+

Mod. FGM

GINA

0.264
0.276
0.293
0.298
0.299

0.268
0.28
0.299
0.304
0.306

0.001
0.01 I
0.023
0.018
0.013

??Experiment
FGM
GINA

25.0

35.0

45.0

55.0

Fiber Volume Fracttan (%)

Fig. 7. Experimental results and analytical predictions for out-of-plane thermal conductivity vs
volume fraction for plain weave E-glass/epoxy composite.

Y. Gowayed and J.-C. Hwang

1184

0.50

W
-GINA

??Experiment
Mod. FGM

0.45

o^
5
$

o&l

i
E

0.35

{
h
L
0

0.30

0
0.25

0.20 L
i !5.()

4j.o

35.0

d.0

Fiber Volume Fraction (%)

Fig. 8. Experimental results and analytical predictions for out-of-plane thermal conductivity vs
volume fraction for plain weave AWepoxy composite.

??Experiment
W
Mod. FGM
-GINA

0.20
25.0

35.0

45.0

-55.0

Fiber Volume Fraction (%)

Fig. 9. Experimental results and analytical predictions for out-of-plane thermal conductivity vs
volume fraction for plain weave Kevlar-49/epoxy composite.

and results from the Graphical Integrated Numerical Analysis (GINA) (Gowayed et al.,
1995) for E-glass/epoxy, AS4/epoxy and Kevlar-49/epoxy test specimens detailed in
section on materials, respectively.
The crimp angles used in the modified FGM and GINA approaches as taken from
the micrographical images for E-glass, AS4 carbon and Kevlar-49 balanced plain weaves
are 2.8, 5.0 and 11.3, respectively. The number of finite element divisions for GINA
is 16 subcells.
Tables l-3 and Figs 7-9 show that the out-of-plane thermal conductivity of plain
weaves increases with the increase of fiber volume fraction. The degree of this increase
could also be attributed to fabric crimp angle, value of fiber thermal conductivities, ratio

Thermal conductivity of plain and 3-D weaves

1185

Table 4. Experimental results and analytical predictions for out-of-plane


thermal conductivity in (W/mC) for 3-D XYZ Toho graphite/epoxy composites
Prediction
Volume fraction
@J)

Experiment

28.77
29.69
38.0

0.3 17 rtr0.024
0.396 + 0.019
0.43 + 0.041

Mod. FGM

GINA

0.267
0.394
0.544

0.274
0.398
0.480

of fiber thermal conductivity along the fiber axis k,,, to that transverse to the fiber axis
k f,t, and ratio of fiber thermal conductivities (kf,,, kf,,) to matrix thermal conductivity.
The AS4/epoxy composite exhibits a pronounced increase in thermal conductivity (57%)
with the increase in fiber volume fraction between 26.8% and 55.7%. This is mainly
attributed to the high kf,, and the 5.0 crimp angle of this fabric. On the other hand,
Kevlar-49/epoxy composites exhibited the least increase in thermal conductivity although
they had the highest crimp angle of 11.3. This is due to the low thermal conductivity in
the transverse fiber direction.
For all practical purposes, GINA and FGM provide a very good estimate for out-ofplane thermal conductivity of the plain weave composite. For E-glass/epoxy and AS4/
epoxy composites, the results are within the experimental error limits. For the Kevlar/
epoxy composite, the percentage of error is around 12%.
A non-linear increase in the thermal conductivity is observed with the increase of
fiber volume fraction of plain weaves. This could be attributed to the high compaction
of fibers at high volume fractions resulting in increased yarn packing factor. Moreover
at high fiber volume fractions, yarns from different fabric layers could touch and form
a heat flow passage reducing the adverse effect of fiber/matrix interface on thermal
conductivity values. Both the modified FGM and GINA do not account for non-linearity
because the micro-level thermal analysis available in literature, and utilized in these
models, do not consider such phenomena. The E-glass/epoxy and the AS4/epoxy plain
weave composites exhibited a higher degree of nonlinearity than the Kevlar-49/epoxy
plain weave composites.
3-D XYZ weaves
Table 4 provides out-of-plane thermal conductivity test results, predictions using the
modified FGM and results from the Graphical Integrated Numerical Analysis (GINA)
(Gowayed et al., 1995) for 3-D XYZ Toho grahite/epoxy weaves. The input for modified
FGM and GINA is taken from the micro-graphical images. The number of finite element
divisions for GINA is 16 subcells.
Results in Table 4 indicate that the out-of-plane thermal conductivity of the graphite/
epoxy composites increases with the increase of the fiber volume fraction in the out-ofplane direction. This is attributed to the high thermal conductivity along the graphite fiber
axis as compared with the thermal conductivity transverse to the fiber direction and the
thermal conductivity of the matrix. Consequently, increasing the volume fraction of
fibers oriented in the out-of-plane direction (V,,) increased the thermal conductivity of the
composite.
As calculated from Table 4, the average error in predictions using GINA is around
8% and the average error for modified FGM predictions for the same samples is 14%.
CONCLUSIONS

In this study, an experimental program is carried out to quantify the out-of-plane


thermal conductivity of three types of balanced plain weaves, E-glass/epoxy, AS4/epoxy
and Kevlar-49/epoxy composites and 3-D XYZ graphite/epoxy woven composites. In the
balanced plain weave composites, a non-linear increase in the out-of-plane thermal
conductivity of the composite is observed with the increase in fiber volume fraction.

1186

Y. Gowayed and J.-C. Hwang

Moreover, the out-of-plane thermal conductivity of the 3-D weaves increased, as


expected, with the increase in the fiber volume fraction in the out-of-plane direction.
The Fabric Geometry Model (FGM), a versatile tool used as a first hand estimate to
predict the stiffness of composite materials, is modified to predict the thermal conductivity of textile composites. Results from the modified FGM are compared with a hybrid
finite element approach Graphical Integrated Numerical Analysis and experimental
analysis.
The modified FGM is able to predict the thermal conductivity of textile composites
with simple preform structures (i.e. plain weaves and 3-D XYZ weaves) with a degree of
accuracy comparable to that of more complex model, such as the Graphical Integrated
Numerical Analysis. Both models are not able to predict the non-linear increase of the
out-of-plane thermal conductivity of the plain weaves with the increase in fiber volume
fraction. The accuracy of the modified FGM is not tested for more complex textile
structures, such as 3-D braids or 3-D angle interlock weaves.
Acknowfedgemenls-The
authors would like to thank Pratt Kt Whitney, NASA Lewis and the National Textile
Center for partial funding of this research activity. Also, they would like to thank Feng Chia University of
Taiwan for providing the 3-D XYZ fabrics.
REFERENCES
ASTM (1985). Standard test method for steady state heat flux measurements and thermal transmission
properties by means of the heat flow meter apparatus. ASTM.
Behrens, E. (1968). Thermal conductivity of composite materials. J. Compos. Mater. 2, 2-17.
Brennan, J., Benstsen, L. and Hasselman, D. (1982). Determination of the thermal conductivity and
diffusitivity of thin fibers by the composite method. J. Mater. Sci 17, 2337-42.
Bruggeman, D. (1935). Dielectric constant and conductivity of mixtures of isotropic materials. Ann. Pays. 24,
636.
Dasgupta, A. and Agrawal, R. (1992). Orthotropic thermal conductivity of plain-weave fabric composites using
a homogenization technique. J. Compos. Mater. 26, 2736-58.
Donea, J. (1972). Thermal conductivities based on variational principals. J. Compos. Mater. 6, 262-66.
Gowayed, Y. A. (1992). An integrated approach to the mechanical and geometrical modeling of textile structural
composites. PhD thesis, North Carolina State University, Raleigh, NC.
Gowayed, Y., Hwang, J. and Chapman, D. (1995). Thermal conductivity of textile composites with arbitrary
preform structures. J. Compos. Tech. Res. 17(l), 56-62.
Han, L. and Cosner, A. (1981). Effective thermal conductivities of fibrous composites. J. Heat Transfer 103,
387-92.
Hashin, Z. (1979). Analysis of properties of fiber composites with anisotropic constituents. J. Appl. Mech. 46,
543-50.
Hasselman, D., Johnson, L., Syed, R. and Taylor, M. (1987). Heat conduction characteristics of a carbon-fiberreinforced lithia-aluminum-silicate glass ceramic. J. Mater. Sci. 20, 2337-42.
Havis, C., Peterson, G. and Fletcher, L. (1989). Predicting the thermal conductivity and temperature distribution in aligned fiber composites. J. Thermophysics 3, 416-22.
James, B., Wostenholm, G., Keen, G. and McIvor, S. (1987). Prediction and measurement of the thermal
conductivity of composite materials. J. Phys.: D, Appl. Phys. 20, 261-68.
Kawabata, S. (1988). Measurements of anisotropic mechanical property and thermal conductivity of single
fiber for several high performance fibers. 4th Japan-US Conference on Composite Materials. June 1988,
Washington DC.
Kregers, A. F. and Melbardis, Y. G. (1978). Determination of the deformability of three-dimensionally reinforced composites by the stiffness averaging method. Polymer Mechanics 1, 3-8.
Lubin, G. (ed.) (1981). Handbook of Composites. Van Nostrand Reinhold, New York.
Maxwell, T. (1904). A Treatise on Electricity and Magnetism, Vol. 1, p. 440. Oxford University Press.
Pastore, C. and Gowayed, Y. (1994). A self consistent fabric geometry model: Modification and application of
a fabric geometry model to predict the elastic properties of textile composites. J. Compos. Tech. Res. 16,
32-36.
Rayleigh, L. (1894). On the influence of obstacles arranged in rectangular order upon properties of the medium.
Philosophical Magazine 34, 48 l-502.
Springer, G. and Tsai, S. (1967). Thermal conductivity of unidirectional materials. J. Compos. Mater. 1,
166-73.
Stuart, M. (ed.) (1990). International Encyclopedia of Composites, Vol. 1, pp. 158-225. VCH Publishers.
Tawil, H., Bensten, L., Bakaran, S. and Hasselman, D. (1985). Thermal diffusivitv of chemicallv vaoor
deposited silicon carbide reinforced with silicon carbide orcarbon fibers. J. Mater. ki. 20, 2337-42. or
Tzadka, U. and Schulgasser, K. (1983). Effective properties of fiber reinforced materials. Transactions of
ASME, J. Appt. Mech. 50, 828-834.
Whyte, D. W. (1986). On the structure and properties of 3-D braided composites. PhD thesis, Drexel University,
Philadelphia, PA.

Você também pode gostar