Вы находитесь на странице: 1из 11

COMPOSITES

SCIENCE AND
TECHNOLOGY
Composites Science and Technology 65 (2005) 18401850
www.elsevier.com/locate/compscitech

Specic properties and fracture toughness of syntactic foam: Eect


of foam microstructures
Erwin M. Wouterson a, Freddy Y.C. Boey a, Xiao Hu a, Shing-Chung Wong
a

b,*

School of Materials Science and Engineering, Nanyang Technological University, Nanyang Avenue, Singapore 639798, Singapore
b
Department of Mechanical Engineering, University of Akron, Akron, OH 44325-3903, USA
Received 14 March 2005; accepted 23 March 2005
Available online 28 April 2005

Abstract
Studies were performed on the specic strength, moduli and fracture toughness of varied microstructures of syntactic foam. The
dierent microstructures were created by using three dierent types of microspheres, namely 3M Scotchlite K15 and K46 glass bubbles, and Phenoset BJO-093 hollow phenolic microspheres, and by changing the volume fractions of microspheres from 0 to
50 vol%. Tension, compression, exural and fracture tests were performed. Results showed that the tensile and exural strengths
decreased with increasing ller content. The behavior of the tensile and exural strength was not aected by the component microspheres. Interestingly, the tensile and exural moduli showed dierent trends for each type of microspheres with increasing ller
content. Results of the compression tests revealed superior behavior of the high density microspheres. The specic fracture toughness data yielded maximum values at 30 vol% for each type of microspheres investigated. Scanning electron microscope studies were
performed to determine the failure mode for each loading condition.
 2005 Elsevier Ltd. All rights reserved.
TM

Keywords: Syntactic foam; Mechanical properties; Fracture toughness

1. Introduction
Syntactic foam is a ternary material system made in
a mechanical way by mixing hollow particles (the ller)
with a resin system (the binder). The hollow particles
may be made of polymer, ceramic, carbon, or metal.
Most often thermoset resins are used as the binder.
Dispersion of the hollow particles creates a porous
material with closed cells. By changing the amount of
hollow ller particles, dierent densities and thus
microstructures of syntactic foam can be created. Syntactic foams are known to possess low density, high
stiness, excellent compressive and hydrostatic
strength, and good impact behavior [16]. Unlike most
*

Corresponding author. Tel.: +1 330 972 8275; fax: +1 330 972


6027.
E-mail address: swong@uakron.edu (S.-C. Wong).
0266-3538/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2005.03.012

other foams, syntactic foam is a material whose density


before curing is the same as that after curing. Such predictability is advantageous in the manufacturing process in aerospace structures.
Using hollow particles, having a lower density compared to the binder material, allows for the manufacturing of light-weight materials with the increase of the
ller content. This type of syntactic foam with a ller
density that is lower compared to the binder can be considered as a special type of particulate-lled polymer
composite (PFPC). Generally, the weight of a PFPC increases with increasing ller content as solid ller particles are most often used. The mechanical and fracture
behaviors of the PFPC were studied extensively [7].
The moduli and fracture properties often improve with
increased solid ller content, given an intrinsically brittle
matrix system and good interfacial bonding between the
ller and the matrix.

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

Most studies on the mechanical and fracture properties of syntactic foams are based on the maximum ller
content of microspheres as this elicits the lowest possible
weight of the composites [13]. This study investigates
the inuence of dierent compositions of syntactic foam
on its specic mechanical and fracture properties. A
comprehensive understanding of the structureproperty
relationship is lacking. Dierent compositions will be
created by varying the type and volume fraction of
microspheres. This route is dierent from that reported
by Gupta et al. [4] whereby dierent microstructures
were created by using dierent types of microspheres
with similar size distributions. However, only the volume fraction of microspheres was kept constant for each
type of microspheres. Bunn and Mottram [5] also reported on the eect of the volume fraction on phenolic
microspheres on the compressive properties. The present
paper compares the results derived from phenolic microspheres to those made of glass shells. It is expected the
interfacial adhesion between phenolic microspheres
and polymers could be stronger.
It is essential that specic properties be discussed in
this paper because the density of syntactic foam changes
with the amount of microspheres introduced. Specic
properties allow for the comparison of the performance
of syntactic foams to other potential foam materials in
sandwich composites such as PU, PVC and aluminum
foams. Furthermore, specic properties aid the comparison between the dierent microspheres used in this
study. By normalizing mechanical and fracture properties against the density, it is believed the results would
be fruitful for guiding future design for syntactic foams
based on specic properties. Results from tension, compression, exure and fracture tests will be discussed in
relation to their microstructures.

2. Experimental work
2.1. Materials and equipment
The syntactic foams for this research were produced
by mechanical dispersion of hollow microspheres in
epoxy resin. Three dierent types of hollow microspheres, namely 3M Scotchlite Glass Bubbles K15
and K46 and Phenoset BJO-093 phenolic microspheres
TM

1841

were used for the ller. Properties of the dierent types


of microspheres are listed in Table 1. The values for the
mean diameter were obtained with a Fritsch particle size
analyzer. The values presented are an average of three
measurements. The values for the average wall thickness
were calculated based on the true density of the microsphere, the density of the microsphere wall material
and the mean sphere diameter. The two types of hollow
glass microspheres were chosen to study the possible differences in mechanical behavior between low and high
density hollow glass microspheres with an increasing
volume fraction. The phenolic microspheres were also
investigated to examine any eects arising from the different nature and interfacial adhesion between ller and
binder. Although the microspheres exhibit dierent
material parameters, it is reasonable to examine the specic properties, which can be quantied and compared
readily. In this paper, we address other material parameters, which form complex interactions with the mechanisms of failure and fracture, qualitatively.
For all the specimens Epicote 1006 epoxy resin was
used as the binder. Epicote 1006 is a combination of liquid bisphenol-A, epichlorohydrin epoxide resin, amine
and polymeric additives. The microspheres were added
to the epoxy while slowly stirring the mixture to minimize gas bubbles in the resin. The microspheres were
added in multiple steps to the epoxy resin to avoid
agglomeration. Due to the low density of the microspheres compared to the binder, the microspheres
showed a tendency to oat to the top surface. This eect
was minimized by stirring the mixtures close to the gel
time of the resin, which was about 60 min for the epoxy.
Scanning electron microscope (SEM) photomicrographs
conrmed the homogeneity of the syntactic foam. After
dispersion, the syntactic foam was compression molded
using an aluminum mold coated with a silicone release
agent. The syntactic foam was left under the press at a
pressure of 1.6 MPa for 1822 h to cure at room temperature. By adding dierent amounts of microspheres to
the matrix, syntactic foams with various densities were
thus created. Fig. 1 illustrates the general microstructure
of syntactic foam.
Fig. 2 shows the measured densities of syntactic
foams with increasing hollow microsphere content.
The measured densities presented in Fig. 2 are an average of about 30 samples. The density was measured by

Table 1
Physical properties of the studied microspheres
Type of microsphere

True density (g/cc)

Static pressure (MPa)

Mean diameter (lm)

Average wall thickness (lm)

Thickness-to-radius ratio

BJO-093
K15
K46

0.25
0.15
0.46

3.44
2.07
41.37

71.5
70a
43.6

1.84b
0.70
1.37

0.052
0.02
0.063

a
b

From [17].
From SEM measurement.

1842

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

2.2. Mechanical tests

Fig. 1. General microstructures under SEM for syntactic foam.

1.25

Theoretical density

Density (g/cm3)

Measured density
1.00

0.75

0.50
0

10

20

30

40

50

vol% microspheres
Fig. 2. Density variation with microsphere content for syntactic foam
containing (j) K15, (m) K46 and (d) phenolic microspheres.

dividing the volume of a specimen by its weight. Fig. 2


shows that all three types of microspheres show a linear
decreasing trend in the density with increasing ller content. The linear trend is expected from the rule of mixtures. The theoretical density of each composition, see
Fig. 2, is calculated according to the rule of mixtures.
The measured density of syntactic foam is lower compared to the theoretical density. The lower value in density is caused by the presence of voids in the composite
which are created during mechanical mixing of the components of syntactic foam. The voids will aect the
mechanical properties. The void content for the dierent
microstructures has been estimated by the following
equation according to ASTM D-2734 [8]:
V v 100  qtheoretical  qmeasured =qtheoretical ;

where Vv is the void volume fraction, qtheoretical is the


theoretical density, and qmeasured is the measured density. The void volume fraction is provided in Table 2.

Three types of mechanical tests, namely tension, atwise compression and exure were performed with different microstructures of syntactic foam. For the
tensile tests, the syntactic foam was machined into a
standard dog-bone bar by means of a TensilKut machine. The specimens were loaded at an ambient temperature using an Instron Model 5567 at a cross-head speed
of 5 mm/min. For each test the tensile strain was recorded with a clip-on strain gauge. The Youngs modulus, Et, was measured from the initial region of
deformation. The results are based on an average of ve
tests. The error bar is the standard deviation for ve
measured values.
Andrews et al. [9] report about the eect of the
specimen size on the mechanical properties. According
to Andrews et al., for closed-cell aluminum foam having a cell diameter of 23 mm, the length of a tensile
specimen should be at least six times the cell diameter
to avoid the specimen dimensions from aecting the
mechanical properties. Because of the small cell size
of syntactic foams, see Fig. 1, all specimens comply
with the requirements set by Andrews et al. and the
specimen sizes used do not aect the mechanical
properties.
For the atwise compression test, syntactic foam was
machined to blocks of 25 25 12 mm3. The length and
width were chosen according to ASTM C365-00 [10].
Due to the short of supply for K46 microspheres, compression tests for K46 specimens were conducted from 0
to 30 vol% microspheres only. In our future work, the
syntactic foam would be used as the core material in
sandwich composites. The height of the specimen was
chosen based on earlier studies [1]. Low aspect ratio
specimens were used to minimize the eect of the shear
stress. The tests were carried out at room temperature
using an Instron Model 4206 with a maximum capacity
of 100 kN. The cross-head speed was 0.5 mm/min. The
atwise compressive yield strength of syntactic foam
was calculated by [10]:
rc

P
;
A

where rc is the compressive yield strength, P the load at


yield, and A is the cross-sectional area.
The atwise compression modulus, Ec, was calculated
by
Ec

mt
;
A

where m is the slope of the initial linear region of the


loaddeection curve, and t is the thickness of the syntactic foam. The results presented are an average of ve
tests. The error bars are derived from the standard
deviation.

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

1843

Table 2
Mechanical properties and void content of the studied compositions of syntactic foam
Tensile
strength
(MPa)

Youngs
modulus
(GPa)

Compression yield
strength
(MPa)

Compression
modulus
(GPa)

Flexure
strength
(MPa)

Flexure
modulus
(GPa)

Fracture
toughness
p
(MPa m)

Epoxy
10% K15
20% K15
30% K15
40% K15
50% K15

29.81
37.32
32.95
23.68
18.96
17.45

2.68
2.76
2.55
2.23
2.07
1.99

85.54
52.95
54.18
44.73
37.96
31.17

1.04
0.88
0.68
0.66
0.63
0.63

78.61
56.61
43.63
27.67
25.59
22.51

2.82
2.31
2.07
1.81
2.06
1.99

0.83
0.95
1.20
1.16
0.94
0.71

2.91
5.14
6.86
9.18
8.85
7.02

10%
20%
30%
40%
50%

K46
K46
K46
K46
K46

43.29
32.87
26.21
23.25
23.23

2.82
3.12
3.29
3.41
3.78

84.61
80.64
76.63
NA
NA

0.95
1.14
1.14
NA
NA

53.32
36.04
31.38
33.39
33.99

2.97
3.13
3.22
3.58
3.86

1.17
1.39
1.27
0.95
NA

4.87
6.07
8.44
10.21
10.02

10%
20%
30%
40%
50%

BJO
BJO
BJO
BJO
BJO

40.84
32.16
25.75
22.29
15.50

2.39
1.88
1.58
1.33
1.17

62.87
51.08
38.11
31.39
25.95

0.80
0.71
0.63
0.57
0.53

60.47
46.70
38.91
31.52
27.22

2.14
1.90
1.52
1.25
1.09

0.87
0.99
1.15
0.92
0.66

3.95
3.69
8.76
8.61
8.03

For the exural tests, syntactic foam was machined to


specimens of 127 12.7 3 mm3 in dimensions. The
tests were performed by an Instron 5567. The span of
the support, S, was chosen to be 48 mm to achieve a
span-to-depth ratio of 16 as recommended by the
ASTM D790-00 [11]. The strain rate was maintained
at 0.01/min. The cross-head speed, z, was calculated
by [11]
z

R  S2
;
6d

where R is the strain rate, S is the span of the support,


and d is the depth of the sample.
The bending modulus, Eb, was calculated by [11]
Eb

S3m
;
4Wt3

where S is the span of the support, and W is the width of


the testing sample.
All specic properties were calculated by normalizing
the measured strength and moduli against the measured
density, q, of the sample.
2.3. Fracture toughness
In the fracture toughness assessment under quasi-static loading, single-edge notched bend (SENB) specimens
were loaded in a three-point bend (3PB) geometry. The
tests were performed by an Instron 5567 at a cross-head
speed of 5 mm/min. Due to the short of supply for K46
microspheres, fracture toughnesses for K46 specimens
were assessed from 0 to 40 vol% microspheres only.
The specimen dimensions were 60 12.7 6.35 mm3.
This specimen geometry satises the requirement for
plane strain conditions [12],


t > 2:5

K Ic
ry

Void
(%)

2
;

where t is the sample thickness, KIc is the critical stress


intensity factor, and ry is the yield strength.
For all the specimens, a constant crack-to-width ratio, a/W, of 0.5 was prepared by a vertical band saw.
A sharp crack was introduced by tapping a fresh razor
blade into a notch.
The critical stress intensity factor, KIc, can be estimated from the following equations [13]:
p
3PS a
;
7
K Ic Y
2tW 2
a
 a 2
 a 3
Y 1:93  3:07
 25:11
14:53
W
W
W
 a 4
25:80
;
8
W
where Y is a shape factor, P is the peak load at the onset
of crack growth in a linear elastic fracture, and a is the
crack length.
After testing, the fracture surface was cut from the
specimen. A thin layer of gold was sputter coated onto
the fracture surface by means of a gold coater (SPI module). The fracture behavior of syntactic foam was then
characterized by means of a Jeol JSM 5410LV, low vacuum SEM.

3. Results and discussion


3.1. Tensile test
Syntactic foam behaves like a linear-elastic material
up to failure when loaded in tension. The material

1844

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

experiences catastrophic failure across a plane perpendicular to the tensile axis. Table 2 presents mechanical
properties and void content of the dierent compositions studied. The trends in the tensile strength, rt, with
increasing microsphere content are rather similar for the
three types of microspheres, except for 50 vol% phenolic
microspheres. Luxmoore and Owen [14] concluded that
a crack will initiate from an oversized void when a composite is subjected to tensile loading. Indeed, numerous
specimens fractured at a cross-section containing a void.
The specic tensile strength, rt/q, for various compositions of syntactic foam are shown in Fig. 3(a). Fig. 3(a)
shows that rt/q increases upon inclusion of a small
amount of microspheres when compared to neat epoxy
resin. The presented data for neat epoxy are obtained
experimentally. However, beyond 10 vol%, a decreasing
trend in rt/q is observed for all types of microspheres.
The decreasing trend in rt/q with increasing ller content indicates that the relative reduction in strength is
larger than the relative reduction in density. Indeed, a
decrease in rt with increasing ller content is observed,
see Table 2. The comparable trends and values suggest
that rt/q is independent of the microsphere type and
size, but only varies with the microsphere volume fraction. The following curve-tting relation between the
rt/q and the volume fraction of microspheres, x, has
been derived for 0.1 < x < 0.5:
rt =q 2157x4 2712x3  1050x2 93x 40:768:

50

40

30

20

10

0
0

(a)

-1

Specific Young's Modulus (GPa.cm .g )

Specific Tensile Strength (MPa.cm3.g-1)

9
Nevertheless, clear advantages regarding other
mechanical properties arise upon introduction of hollow
microsphere as shown in the discussion that follows.
Luxmoore and Owen suggested [14] that the failure of
the foam is attributed to the failure of the resin matrix.
The non-linearity in rt/q is caused by the reduction
in the area of the epoxy matrix in a cross-sectional area.

Introduction of hollow microspheres reduces the epoxy


volume fraction and increases the inhomogeneity content, consequently reducing the tensile strength of the
composite as a result of poor interfacial strength between the matrix and the ller. The authors assume that
the matrix serves as the load-bearing phase in the composite whereas the hollow microspheres only provide
light weight and minimal strengthening eect. The
reduction in load-bearing volume outweighs the increase
in sti microsphere shells. Around 40 vol% of ller content, a minimum is observed for all the microspheres.
We believe that 40 vol% of microspheres is to be the
maximum amount of ller which can be fully wetted
by the epoxy matrix in this system. The change in properties and behavior of syntactic foam around 40 vol%
was also observed by Bunn and Mottram [5]. Nevertheless, clear advantages regarding other mechanical properties arise upon introduction of hollow microsphere as
shown in the discussion that follows.
It has been reported that the Youngs modulus, Et,
generally decreases with increasing content of hollow
microspheres [15,16]. Only Bardella and Genna [17] report an increasing trend in Et for syntactic foams containing K37 (q = 0.37 g cm3) microspheres. The
results from the current research, as presented in Fig.
3(b) shows similarity to those in [17] in the sense that
the trend for the specic Youngs modulus, Et/q, with
increasing ller content depends on the type of microspheres used. K46 microspheres show a rather linear increase in Et/q, whereas K15 show a constant trend.
Based on the data presented in Table 2, it is known that
the constant trend in Et/q is caused by a proportionate
decrease in the Youngs modulus and the density. For
hollow phenolic microspheres, a decrease in Et with
increasing ller content is observed. The value obtained
for neat epoxy resin is too low considering the trends in
Et for syntactic foams containing the three dierent

10

20

30

vol % microspheres

40

50

0
0

(b)

10

20

30

40

50

vol % microspheres

Fig. 3. Specic tensile properties of syntactic foam containing (j) K15, (m) K46 and (d) phenolic microspheres: (a) specic tensile strength; (b)
specic Youngs modulus.

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

types of microspheres. The lower value for neat epoxy


resin could be caused by the presence of voids. The differences between the results for K46 and K15 microspheres can be attributed to their size and thickness. It
has been shown by several researchers that, in general,
for microspheres of the same material composition, improved mechanical properties are obtained for the
microspheres with a higher density. Higher density for
microspheres of the same material is often associated
with a larger thickness-to-radius ratio, t/R, see Table
1. Syntactic foam containing microspheres of the same
material having a larger t/R often exhibit improved
mechanical properties [18]. The dierence between the
phenolic and K15 glass microspheres explains that apart
from microsphere size, the material composition of the
microsphere is of equal importance for the mechanical
properties. Phenolformaldehyde has an Et of about
6.8 GPa whereas soda lime glass has an Et of about 77
GPa. The dierence in Et is believed to be an important
factor to the dierence in the trends of hollow glass and
phenolic microspheres. This relationship between the
material of the matrix and the inclusion was studied
by Pawlak and Galeski [19]. It was found that the higher
the Youngs modulus of the inclusion compared to the
epoxy used, the higher the stresses developed in the
material. Further they observed that the position of
the maximum stress at the spherical inclusion changed
from the pole for hard inclusion to the equator for soft
inclusion.
Fig. 4 shows an SEM fractograph of a tensile specimen; the micrograph conrms the role of the binder in
the failure of syntactic foam under tensile loading. The
micrograph shows a rough surface of the epoxy resin
after brittle fracture. Numerous step structures can be
identied in Fig. 4 which indicates the plastic yielding
of the epoxy resin. Besides plastic yielding, intact microspheres are observed on the fracture surface. The fact

1845

that the crack has grown over the interface between


the matrix and the microsphere indicates the presence
of debonding. Debonding of hard inclusions was also reported by Pawlak and Galeski [19]. The debonding is
caused by the complex stress state around the particle
matrix interface which will not be further discussed in
this paper.
3.2. Compression test
Fig. 5 shows the stressstrain curves of atwise compression tests of syntactic foams containing various
amounts of K15 microspheres. The compression curves
of syntactic foam were previously discussed by other
investigators [1,2,5]. For each individual curve, three different regions can be identied in Fig. 5. The rst region
(I) is characterized by an almost linear-elastic behavior
of the syntactic foam. The region ends when the material
starts to yield and reaches its compressive yield strength.
Upon yielding, the microspheres become crushed under
the compression load and severe damage occurs. Yielding and inelastic damage occur during the test as the
sides of the specimen barrel outward under compression
loading. Barreling of the specimen is caused by friction
between the contact surfaces. It aects the stress of the
syntactic foam under loading. The second region (II)
of the atwise compression curves in Fig. 5 is characterized by relatively horizontal plateaus. The horizontal
plateau is attributed to the implosion of the hollow
microspheres under the increasing compression load.
Syntactic foams with higher fractions of hollow microspheres show a larger horizontal plateau and thus a larger strain. The third region (III) is characterized by a
steep increase in the loaddisplacement curve. The steep
increase is caused by a large number of microspheres
being crushed and compacted, and the maximum den-

Compressive Strength, c (MPa)

160
140
120
100

III

80

60

II

40
20

increasing volume fraction

0
-20
-0.1

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Engineering strain (mm/mm)


Fig. 4. Fracture surface of a syntactic foam specimen perpendicular to
the direction of applied tensile loading.

Fig. 5. Compression stressstrain curves of syntactic foam with


various amounts of K15 microspheres content. Regions IIII are
sequentially observed and indicated.

1846

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

(a)

3.3. Flexure testing of syntactic foam


Figs. 7(a) and (b) show typical stressstrain curves for
K46 and BJO-093 microspheres, respectively. K15
microspheres show a similar exure behavior as BJO093 microspheres. For all specimens, it is observed that
the strain is reduced with increasing ller content. Especially, the syntactic foams containing K46 microspheres
show a larger reduction in failure strain. The larger
reduction in strain for K46 microspheres is attributed
to the higher strength of the K46 microspheres, see Table 1. Fewer microspheres will fracture under the applied load which prevents stress relief in the material
[20]. On the contrary BJO-093 and K15 microspheres
fracture more easily, relieving the stress in the material,
and thus showing a larger strain.
Upon fracture some plastic deformation is observed
for most compositions of syntactic foam with up to
30 vol% of microspheres. The amount of plastic deformation decreases with increasing ller content since
the plastic deformation is attributed to the behavior of
the epoxy binder. Syntactic foams containing hollow
phenolic microspheres show larger plastic deformation
compared to K46 and K15 microspheres. The latter
might be attributed to ductile deformation of hollow
phenolic microspheres as shown in Fig. 8; an observation that has not been seen for hollow glass microspheres. The plastic deformation of syntactic foam
containing hollow phenolic microspheres can be clearly
seen on the fracture surface of specimens due to stress
whitening. The stress whitening cannot be distinguished
on syntactic foam containing hollow glass microspheres
since the resulting syntactic foam is white in color.
During exure loading, the specimen is subjected to
compressive stresses on the top part of the specimen
and to tensile stresses at the lower part of the specimen.
From the tension and compression data presented in the

1.50

-1

Specific Compression Modulus (GPa.cm .g )

100

-1

Specific Compression Strength (MPa.cm .g )

sity is being reached. The point, where region III starts,


is considered to be point of failure for the syntactic foam
as this is the point where most of the load bearing microspheres have been crushed.
The overall results of the atwise compression tests
are given in Fig. 6 and the void content is summarized
in Table 2. Fig. 6(a) shows that the specic compressive
yield strength, ryc/q, decreases with increasing ller content for phenolic and K15 microspheres. The microspheres act as voids and weaken the structure.
Nevertheless, an upward trend in ryc/q is observed for
K46 microspheres. The upward trend is attributed to a
relatively minor decrease in the compressive yield
strength, see Table 2, compared to the decrease in density. From the dierence in trends between the K15
and K46 microspheres, it is induced that the specic
compressive yield strength is inuenced by the wall
thickness of microspheres.
The results for the specic compressive modulus, Ec/q,
are shown in Fig. 6(b). Again K46 microspheres perform better as shown. The performance of K46 microspheres is also attributed to their t/R ratio. For all
microspheres investigated, it appears a densication
process takes place at low ller content (10 vol%).
The densication is caused by a compaction process of
the matrix polymer in the presence of voids in curing.
As microsphere content increases and matrix volume decreases, the microsphere takes up more load under compression and, as a result, the specic compressive moduli
increase. The trend under compression is in contrast to
that under tension. K15 and phenolic microspheres
show a rather similar behavior in Ec/q. Both show a
minimum around 20 vol%, after which the Ec/q slightly
increases. This increase is higher for K15 microspheres.
The specic compressive stiness levels o as the increase in stiness is counter-balanced by the increases
in density of the composite.

80

60

40

20

0
0

10

20

30

40

vol% microspheres

50

60

(b)

1.25

1.00

0.75

0.50

0.25

0.00
0

10

20

30

40

50

60

vol% microspheres

Fig. 6. Specic compressive properties of syntactic foam containing (j) K15, (m) K46 and (d) phenolic microspheres: (a) specic compressive yield
strength; (b) specic compressive modulus. Due to the short of supply for K46 microspheres, compression tests for K46 specimens were conducted
from 0 to 30 vol% microspheres only.

80

80

60

60

Stress (MPa)

Stress (MPa)

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

40

20

1847

40

20
Increasing volume fraction

0
0.00

(a)

0.01

0.02

0.03

0.04

0.05

0.06

Strain (mm/mm)

Increasing volume fraction

0.00

(b)

0.01

0.02

0.03

0.04

0.05

0.06

Strain (mm/mm)

Fig. 7. Flexure stress strain curves of syntactic foam containing various amounts of (a) K46 and (b) phenolic microspheres. The arrow indicates the
direction of increasing ller content.

Fig. 8. A deformed hollow phenolic microsphere.

previous paragraphs, it is expected that syntactic foam


subjected to exure loading will fail at the side under
tensile stresses. Indeed, failure at the lower specimen
side under tensile stresses is observed during experimental work. All specimens fail at the center of the support
span. No indentation of the indenter is observed. The
comparable failure mode under tensile loading and
three-point-bending explains why there is some resemblance between Figs. 3 and 9. The only dierence
observed is the behavior of syntactic foams with
010 vol% of microspheres.
For the specic exural strength, rf/q, see Fig. 9(a),
some dierences between the dierent compositions are
observed. The dierences are mainly caused by the difference in density between the dierent compositions
as the values and trends are rather similar for the maximum exural strengths, see Table 2. All types of microspheres show a decreasing trend in the maximum specic
exural strength with increasing ller content. Similar to
the tensile test results, the specic strength approaches a

minimum around 4050 vol% of ller content. Based on


the data presented in Table 2, it can be said that the horizontal trend for the hollow phenolic and K15 microspheres at high volume fractions, is caused by a
proportionate decrease in the exure strength and
density.
The trends and values in the specic exural modulus, Ef/q, see Fig. 9(b), are consistent with the results
presented in Fig. 3(b). The similarity is again attributed
to the tensile failure mode of syntactic foam under three
point bending. The K46 microspheres show an increase
in Ef/q with increasing ller content. K15 microspheres
show a constant value for Ef/q whereas phenolic microspheres show a decrease with increasing ller content.
The constant value in Ef/q from K15 microspheres is
attributed to a similar relative decrease in the exure
modulus and density, see Table 2.
SEM fractographs of the specimens tested under
three point bending reveal the dierent loading modes
at the upper and lower parts of the cross-sectional
area of the specimen. Fig. 10(a) shows the fracture
surface of the lower part of the cross-section, which
is loaded in tension. The fractograph is consistent with
that in Fig. 4. The surface is characterized by a rough
surface. In between the tension and compression sides
of the specimen, an area with relative smooth features
is observed. The transition between the tensile surface
and the compressive surface is shown in the right side
of Fig. 10(a). A typical fracture surface of the upper
part of the cross-section, loaded in compression, is
shown in Fig. 10(b). The compressive surface shows
step structures behind the microspheres which we previously observed [21]. The step structures diminish
with increasing ller content and are less pronounced
for phenolic microspheres. Besides the step structures,
debonded microspheres can be identied. Debonded
microspheres indicate a poor interface between the
matrix and the ller.

1848

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850


6
-1
3

70
60
50
40
30
20
10
0

(a)

Specific Flexure Modulus (GPa.cm .g )

-1

Specific Flexure Strength (MPa.cm .g )

80

10

20

30

40

50

vol % microspheres

10

20

30

40

50

vol % microspheres

(b)

Fig. 9. Specic exural properties of syntactic foam containing (j) K15, (m) K46 and (d) phenolic microspheres: (a) specic exure strength;
(b) specic exure modulus.

Fig. 10. SEM micrograph of the fracture surface of syntactic foam containing 10 vol% K15 microspheres under (a) tension, and (b) compression.
The black arrows denote the direction of crack propagation.

0.5

Fig. 11 shows the specic fracture toughness, K1c/q,


for various compositions of syntactic foam. The actual
fracture toughnesses, KIc, together with the void content
are presented in Table 2. The most distinctive features in
Fig. 11 are the signicant increase in K1c/q for ller content up to 30 vol% for all types of microspheres, and the
decrease in K1c/q beyond 30 vol% of ller content. The
maximum in the trend for the fracture toughness has
been reported for other particulate composites
[7,13,22]. The increase in K1c/q for 030 vol% ller content suggests the presence of a toughening mechanism
which increases the fracture energy compared to neat
epoxy resin. Again the hollow glass microspheres outperform the hollow phenolic microspheres with K46
microspheres resulting in the highest value for K1c/q,
suggesting that the wall thickness and density of the
microspheres aect the fracture toughness of the syntactic foam.

-1

Specific Fracture Toughness (MPa.m .cm .g )

3.4. Fracture toughness of syntactic foam


1.5

1.0

0.5

0.0
0

10

20

30

40

50

vol % microspheres
Fig. 11. Specic fracture toughness vs. ller content for syntactic foam
containing (j) K15, (m) K46 and (d) phenolic microspheres. Due to
the short of supply for K46 microspheres, fracture toughnesses for K46
specimens were assessed from 0 to 40 vol% microspheres only.

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

1849

Fig. 12. SEM micrographs of the fractured surface of SENB specimens of syntactic foam containing 10 vol% microspheres: (a) K15 microspheres;
(b) hollow phenolic microspheres.

We believe that the type of major toughening mechanism has changed at the transition point of 30 vol% of
microsphere content. The toughening from 0 to
30 vol% microspheres is inuenced by a combination
of the ller stiening eect and crack bowing mechanisms in the presence of round-shaped llers. The
change in toughening mechanism is conjectured to be
caused by the inter-particle spacing theory as discussed
by Lee and Yee [13] for glass bead/epoxy composites.
Increase in microsphere content will decrease interparticle separation between microspheres. The increase
of microsphere content beyond the complete wetting
ability of epoxy also introduces inter-sphere sliding
and stress concentration. Higher volume fractions of
microspheres allow more microspheres to debond from
the matrix. Debonding is accompanied by premature
cracks. If the direction of these cracks is parallel to the
crack growth direction, the subcritical cracks act as precursors and facilitate crack propagation.
SEM was used to elucidate the toughening mechanisms in the various microstructures of syntactic foam,
see Figs. 12 and 13. Fig. 12 shows the fracture surface
of syntactic foams containing low volume fractions of

microspheres, where Fig. 13 shows the fracture surface


of syntactic foams containing high volume fractions of
microspheres. Clearly, Figs. 12(a) and (b) show that step
structures prevail for the microstructures containing low
volume fractions of microspheres. These step structures
were previously observed [21]. These step structures are
considered to be the evidence of the existence of the
crack front bowing mechanism. This mechanism was
rst observed experimentally by Lange in 1970 [23].
Lange suggested that an approaching crack front is pinned by a rigid particle. Secondary crack fronts will be
formed and these secondary crack fronts will bend/
bow between the rigid particles. There will be a point
when the crack front breaks away from the rigid particle. At this point, the arms of the secondary crack fronts
will come together and form a characteristic step structure as both secondary crack fronts propagate at a different crack plane. These characteristic step structures
are also called tails or lances. Many studies on the
fracture toughness of particulate composites consider
the crack bowing as the main toughening mechanism.
As hollow glass microspheres exhibit higher fracture
toughness compared to hollow phenolic microspheres,

Fig. 13. SEM micrograph of the fractured surface of SENB specimens of syntactic foam containing large amounts of microspheres: (a) 40 vol% K46
microspheres; (b) 50 vol% hollow phenolic microspheres.

1850

E.M. Wouterson et al. / Composites Science and Technology 65 (2005) 18401850

with K46 microspheres outperforming K15 microspheres, it is surmised that the thickness-to-radius ratio
and the wall material of the microsphere aect the size
of the step structure and thus K1c/q.
Figs. 13(a) and (b) show the microstructures of syntactic foam containing 40 vol% K46 and 50 vol% hollow
phenolic microspheres, respectively. Compared to Fig.
12, step structures are absent in Fig. 13. Instead, the
dominant fracture mechanism that is observed is debonding of microspheres. The existence of excessive debonding introduces stress concentration and premature
cracks leading to reduced fracture toughness.
4. Conclusions
From the results presented in this paper, it can be
concluded that the specic properties of syntactic foam
depend on the types and volume fractions of microspheres utilized in the syntactic foam. Increase in microsphere density (K46 vs. K15) and the thickness-to-radius
ratio led to an increase in specic tensile stiness. The results for the tensile and exural tests were comparable
due to the fact that both types of tests exhibited the
same failure mode. Both tests elicited a decreasing trend
in specic strength with increasing ller content. A distinct trend in compressive behavior was noted in contrast to the tensile failure. The compression tests
revealed the excellent compressive properties of syntactic foam and in particular the superior performance of
K46 microspheres, giving rise to higher compressive
yield strengths and moduli compared to K15 and phenolic microspheres.
From the fracture toughness tests, it was concluded
that all types of studied microspheres show a similar
trend in the specic fracture toughness with increasing
ller content. For lower ller content an increase in
the specic fracture toughness was observed. The increase reached a maximum after which a decrease in
the specic fracture toughness was seen. The change in
behavior was attributed to a change in the dominant
toughening mechanisms from ller stiening, crack
front bowing to excessive debonding of microspheres
in reduced matrix volume. This work, however, demonstrated the usefulness of a combination of desired properties for syntactic foam such as light-weight high
stiness, high compression and high toughness.
Acknowledgments
The authors thank Nanyang Technological University and DSO National Laboratories Singapore for the
support of this work. One of us (SCW) acknowledges
the support of NSF Grant# CMS 0335390 administered
by the Mechanics and Materials Program.

References
[1] Gupta N, Kishore, Woldesenbet E, Sankaran S. Studies on
compressive failure features in syntactic foam material. J Mater
Sci 2001;36(18):448591.
[2] Gupta N, Woldesenbet E, Kishore. Compressive fracture features
of syntactic foams microscopic examination. J Mater Sci
2002;37(15):3199209.
[3] Kim HS, Oh HH. Manufacturing and impact behavior of
syntactic foam. J Appl Polym Sci 2000;76(8):13248.
[4] Gupta N, Woldesenbet E, Mensah P. Compression properties of
syntactic foams: eect of cenosphere radius ratio and specimen
aspect ratio. Compos Part A 2004;35A(1):10311.
[5] Bunn P, Mottram JT. Manufacture and compression properties
of syntactic foams. Composites 1993;24(7):56571.
[6] Shutov FA. Syntactic polymer foams. In: Klempner D, Frisch
KC, editors. Handbook of polymeric foams and foam technology. Hanser Publishers; 1991. p. 35574.
[7] Rothon RN. Particulate-lled polymer composites. New
York: Longman, Scientic and Technical; 1995.
[8] Anon. Standard test methods for void content of reinforced
plastics. ASTM D 2734-94. New York: ASTM; 1994.
[9] Andrews EW, Gioux G, Onck P, Gibson LJ. Size eects in ductile
cellular solids. Part 2: Experimental results. Int J Mech Sci
2001;43:70113.
[10] Anon. Standard test method for atwise compressive properties
of sandwich cores. ASTM C 365-00. New York: ASTM; 2000.
[11] Anon. Standard test method for exural properties of unreinforced and reinforced plastics and electrical insulating materials.
ASTM D 790-00. New York: ASTM; 2000.
[12] Hertzberg RW. Deformation and fracture mechanics of engineering materials. New York: Wiley; 1996.
[13] Lee J, Yee AF. Fracture of glass bead/epoxy composites: on
micro-mechanical deformations. Polymer 2000;41(23):836373.
[14] Luxmoore AR, Owen DRJ. The mechanics of syntactic foams. In:
Hillyard NC, editor. The mechanics of cellular plastics. Barking: Applied Science Publishers; 1980. p. 35991.
[15] Krstic VD, Erickson WH. A model for the porosity dependence of
Youngs modulus in brittle solids based on crack opening
displacement. J Mater Sci 1987;22(8):28816.
[16] El-Hadek MA, Tippur HV. Simulation of porosity by microballoon dispersion in epoxy and urethane: mechanical measurements
and models. J Mater Sci 2002;37:164960.
[17] Bardella L, Genna F. On the elastic behavior of syntactic foams.
Int J Solids Struct 2001;38(4041):723560.
[18] Fine T, Sautereau H, Sauvant-Moynot V. Innovative processing
and mechanical properties of high temperature syntactic foams
based on a thermoplastic/thermoset matrix. J Mater Sci
2003;38(12):270916.
[19] Pawlak A, Galeski A. Determination of stresses around beads in
stressed epoxy resin by photoelasticity. J Appl Polym Sci
2002;86(6):143644.
[20] Gupta N, Li G, Jerro D, Woldesenbet E, Pang S-S. Eect of
nano-size clay particles on the exural properties of syntactic
foams. In: Proceedings of the 62nd annual technical conference of
the society of plastics engineers, ANTEC 2004, USA, Society of
Plastics Engineers, p. 132024.
[21] Wouterson EM, Boey FYC, Hu X, Wong S-C. A study on
fracture toughness of syntactic foam. In: International conference
on materials for advanced technologies (ICMAT) proceedings,
MRS Singapore; 2003.
[22] Kawaguchi T, Pearson RA. The eect of particlematrix adhesion
on the mechanical behavior of glass lled epoxies. Part 2. A study
on fracture toughness. Polymer 2003;44(15):423947.
[23] Lange FF. The interaction of a crack front with a second-phase
dispersion. Philos Mag 1970;22(179):98392.

Вам также может понравиться