You are on page 1of 13

International Journal of Fatigue 24 (2002) 847–859

www.elsevier.com/locate/ijfatigue

Environmental fatigue behavior and life prediction of unidirectional


glass–carbon/epoxy hybrid composites
*
Ying Shan, Kin Liao
School of Materials Engineering, Nanyang Technological University, Nanyang 639798, Singapore

Received 15 October 2000; received in revised form 10 August 2001; accepted 20 November 2001

Abstract

Unidirectional glass fiber reinforced and glass–carbon fiber reinforced epoxy matrix composite specimens were subjected to
tension–tension fatigue in air and in distilled water at 25 °C. While no significant change in fatigue life was observed for both
types of specimens tested in air and in water when cyclically tested at 85% of average ultimate tensile strength (UTS), the detrimental
effect of water becomes apparent at lower stress levels of 65 and 45% UTS. Compared to specimens tested in air, cyclic loading
in water results in shorter fatigue lives for both glass and hybrid specimens. While all of the glass fiber specimens did not survive
to 106 cycles when cyclically loaded in water, hybrid specimens (with 25% carbon fiber (by volume), 75% glass fiber (by volume),
30% total fiber volume fraction) showed better retention in structural integrity under environmental fatigue, for fatigue lives up to
107 cycles, a consequence of the corrosion resistant of carbon fiber. Thus it is shown, by incorporating appropriate amount of
carbon fibers in glass fiber composite, a much better performance in fatigue can be achieved for glass–carbon hybrid composite.
A simple life prediction model for the hybrid composite is proposed. Model predictions are compared with experiments results
from both laminated interply and intermingled intraply hybrid composites. Results suggest that synergistic effect of the reinforcing
fibers is critical in governing the fatigue behavior of intraply hybrid composite.  2002 Elsevier Science Ltd. All rights reserved.

Keywords: Hybrid composites; Environmental fatigue; Life prediction

1. Introduction are proposed based on such a ‘constant strain’ assump-


tion, where all of the low elongation fibers in a hybrid
1.1. Hybrid composites composite are assume to break at their failure strain, fol-
lowed by failure of the high elongation (HE) fibers. A
A hybrid composite consists of two or more types of schematic strength diagram constructed from these mod-
reinforcing fibers in one or more types of matrices. By els is illustrated in Fig. 1. Point A denotes the tensile
hybridizing two or more types of fiber in a matrix allows strength of the high elongation fiber composite, and
a closer tailoring of composite properties to satisfy spe- point D that of the low elongation fiber composite. The
cific requirements compared with composites with only strength of the hybrid composites is then given by the
a single type of fiber. Uniaxial tensile behavior of unidi- two straight lines AC and CD. To the right side of point
rectional hybrid composites has been studied extensively C, failure of the low elongation phase leads to immediate
since the mid 1970s. Whilst the elastic modulus in the failure of the entire hybrid composites since there is
fiber direction can be predicted by rule-of-mixture insufficient high elongation fibers to sustain the load
(ROM) approach, the tensile strength of hybrid com- after failure of the low elongation component. The
posites does not obey the ROM since the low elongation strength, shybrid, of the hybrid composite is then given by
(LE) fibers are expected to break first when their failure shybrid ⫽ sLEnfLE ⫹ eLEEHEnfHE (1)
strain is reached. Simple tensile strength models [1,2]
where nf is the volume fraction, s the tensile strength,
e the maximum strain, E, the Young’s modulus, and
* Corresponding author. Tel.: +65-790-6258; fax: +65-790-9081. subscripts LE and HE denote low elongation and high
E-mail address: askliao@ntu.edu.sg (K. Liao). elongation fiber, respectively. To the left side of point

0142-1123/02/$ - see front matter  2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 1 4 2 - 1 1 2 3 ( 0 1 ) 0 0 2 1 0 - 9
848 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

1.2. Environmental fatigue

When a material is cyclically loaded in the presence


of an environmental agent, such as moisture and/or tem-
perature, it is said to be under environmental fatigue.
In general, elevated temperatures and humidity shorten
fatigue lives of fiber reinforced polymer matrix com-
posites [9]. While preconditioning under relatively low
relative humidity may not have a significant effect on
glass fiber reinforced plastics (GFRP) [10], precon-
ditioning in water at elevated temperatures (⬎75 °C)
often has a deleterious effect on the performance of
GFRP [10,11]. The fatigue performance of carbon fiber
Fig. 1. Strength curve constructed by constant strain model and rule-
of-mixture.
reinforced plastics (CFRP) has been shown to be much
better than that of GFRP, even at elevated temperature
and high humidity. For instance, Jones et al. [11] showed
that preconditioning carbon/epoxy composite in boiling
C, there are sufficient high elongation fibers to sustain water for three weeks has practically no effect on the
the load after failure of the low elongation fibers at stress fatigue performance compared to specimens without pre-
levels given by AC. The ultimate strength of the hybrid conditioning, but glass composites did degrade at a much
at this volume fraction is then given by faster rate.
shybrid ⫽ sHEnfHE (2) Most of the previous studies on environmental fatigue
of fiber reinforced plastics (FRP) were focused on vari-
Since the crack arresting mechanism of the high ous preconditioning situations, and data on environmen-
elongation components is not taking into consideration tal fatigue of FRP tested under fluid immersion con-
in the constant strain model, it is an over simplification, ditions were very limited and results sometimes
yielding a lower bound of hybrid strength. Deviation of contradicting. However, it is indicated by those data that
experiment data from this model is regarded as syner- fatigue under fluid immersion is more severe than pre-
getic effect or hybrid effect. conditioning. Sumsion et al. [12] has compared flexural
Generally speaking, the properties of a hybrid com- fatigue behavior of uniaxial (0°) and cross-plied (±45°)
posite are considered to be an average of those of the graphite/epoxy specimens tested in air and in water at
parent composites, and the fatigue responses are often room temperature, substantial reduction in fatigue life
weighed by a ROM relation. Philips [3] reported that for was observed for the latter case. In contrast, Komai et
glass–carbon cloth an increase in tensile fatigue strength al. [13] reported fatigue life of unidirectional CFRP
is proportional to the amount of carbon fibers. Studies tested in immersed deionized water was slightly longer
by Marom et al. [4] on flexural fatigue behavior of unidi- than those tested in the air. They attributed this to the
rectional carbon–aramid hybrid composites showed that enhanced ductility of the polymer matrix caused by
the slope of the S–N curve was given by a ROM water absorption. They also found that specimens sub-
relationship. In a study on the unidirectional carbon– jected to preconditioning in water at various conditions
kevlar epoxy hybrids, Fernando et al. [5] reported the before cyclically loaded in water showed considerable
fatigue strength of the hybrids varies linearly with com- reduction in fatigue life. McBagonluri et al. [14] perfor-
position. Also, they claimed, after examining the fatigue med environmental fatigue on unidirectional pultruded
ratio (fatigue strength divided by tensile strength), that E-glass/vinyl ester composite by loading samples in a
their hybrids exhibited a positive deviation from ROM fluid cell. They reported that after preconditioning in salt
or the actual fatigue strength is higher than the result water at 65 °C for 5 months the fatigue life for speci-
given by ROM predictions. In a subsequence paper mens tested in salt water was considerably reduced com-
devoted to the fatigue behavior of unidirectional glass– pared to that of as-delivered specimens cyclically tested
carbon hybrid composites [6], similar results were in air. However, when normalized to their respective
reported by Dickson et al. In contrast with the amount of UTS, S–N curves revealed the same slope. It was there-
literature on experimental results of hybrid composites, fore concluded that the fatigue damage mechanism was
however, few paper discuss modeling the fatigue the same for both conditions and was fiber dominated.
behavior of hybrid composites. An often debated matter Recently, Liao et al. [15] have developed long-term
is whether it is possible to determine the fatigue response flexural fatigue data for unidirectional pultruded E-
of a hybrid composite from the properties of its parent glass/vinyl ester composite tested in water and NaCl sol-
composites. utions. Under low stress, high cycle fatigue, substantial
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 849

reductions in fatigue life under fluid immersion were The total fiber volume fraction was kept at approxi-
observed. mately 30% for both all-glass and hybrid specimens.
It is indicated by majority of the available data that Hybrid composite panels of relative glass to carbon fiber
severe material degradation could be introduced to the volume fractions of 100:0, 95:5, 90:10, 75:25, 50:50, and
material under environmental fatigue of GFRP, even at 0:100 were fabricated to assess the effect of relative
low stress levels. In order to enhance the durability of glass to carbon volume ratio on the mechanical proper-
GFRP, hybrid composite systems using both carbon and ties of the hybrids. Bar specimens of dimension
glass fibers have been proposed [16,17]. As mentioned 100×12×1.5 mm were cut from the unidirectional panel
previously, hybrid composites seem to be an attractive after being released from the mold, and the two edges
material in resisting environmental attack. Both fatigue of each specimen were polished to avoid edge effects.
performance and environmental resistance of glass–car- Composite print circuit board (PCB) end-tabs of length
bon hybrid composite is enhanced compared to all-glass 25 mm were bonded to the specimen surface, leaving a
composites [9]. Compared to interply hybrid composites gauge length of 50 mm. Quasi-static tensile tests were
constructed by laminating together distinct layers of performed using an Instron 5567 universal test
glass-fiber-reinforced composite and carbon-fiber- machine to determine their ultimate tensile strength
reinforced composite, the intraply hybrid composite, in (UTS), at a cross-head speed of 1 mm/min. At least four
which bundles of carbon and glass fibers are mixed specimens were tested for each group.
within a single layer, is considered more desirable in Fatigue and environmental fatigue were carried out at
terms of reduced mismatches. However, very limited stress levels of 30, 45, 65, and 85% UTS using an
work has been carried out on intraply hybrid systems Instron 8516 servo-hydraulic test frame. Two groups
because they are more difficult to fabricate. of composites specimens were employed to characterize
In this paper, results of a study on fatigue and environ- the fatigue and environmental fatigue behavior: all-glass
mental fatigue behavior of unidirectional glass com- 100:0 and hybrid 75:25. Specimens were cyclically
posite and glass–carbon intraply hybrid composite are loaded using a sinusoidal wave function at 10 Hz for
presented, and thereby it is shown that the hybrid 65% and 85% UTS, and 20 Hz for 30 and 45% UTS,
approach is viable in enhancing long term durability. A at R ⫽ 0.1 (where R is the ratio of minimum to the
simple phenomenological fatigue life prediction model maximum cyclic stress) at 25 °C. Environmental fatigue
for unidirectional hybrid composites is also presented. was performed with specimens immersed in distilled
water, shown schematically in Fig. 2, where the gauge
section of a specimen was wrapped with a rubber bag
2. Materials and experiment filled with water-containing cotton with the lower end
sealed. Fatigue tests were interrupted periodically to
Unidirectional glass/epoxy and intraply glass– measure the tensile modulus of a specimen by means
carbon/epoxy specimens were fabricated by a wet wind- of an extensometer (Instron model 2620-601). Surface
ing and hot pressing method described by Peijis et al. related damage during fatigue was also visually exam-
[18]. E-glass fiber, PAN–HTA carbon fiber, and an Epic- ined. Specimens that have survived to 107 cycles were
ote 600 epoxy, a room temperature cured resin, are used
for this study. Fibers are used as-received without sur-
face treatments. Continuous glass fiber bundles were first
wound in parallel on a rectangular metallic frame. In
order to enhance the environmental resistance, carbon
fibers were wound after and on top of the glass fiber
bundles with an even spread. As a consequence, more
carbon fiber was near the surface of the resulting com-
posite plate. The epoxy matrix was prepared by first mix-
ing with the hardener at a specified ratio recommended
by the vendor, and subsequently degassed in a vacuum
oven for 5 min at a pressure of 25 mm Hg. Fibers were
wetted by pouring the degassed epoxy into a plastic bag
containing the fiber bundle-frame assembly. They were
then put into an open-ended mould pre-pasted with
release film to ensure ease of detachment of the com-
posite after curing. The whole assembly was subjected
to hot press for 2 h at 80 °C and a pressure of 20 t.
When curing was completed, the composite plate was
removed from the mold and post cured at 110 °C for 6 h. Fig. 2. Environmental fatigue specimen setup.
850 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

with some fiber splitting and bundle failure. Grip-related


failures were seen in some cases, which could be a cause
for the observed relatively low tensile failure strain. As
a result of the filament winding process, the fiber roving
may not be perfectly aligned and uniformly distributed
within the matrix. Also, because of the relatively small
size of the test specimens (0.18 cm2 nominal cross sec-
tional area), some specimens may contain more fiber
roving than others, causing scattering in tensile strength
data. Unlike most of the reports on the tensile behavior
of hybrid composites, an obvious drop in strength at low
carbon content is not seen. Based on the predictions of
Fig. 3. Typical stress–strain curves of various hybrid compositions. the constant strain model, the lowest strength should
The curves of hybrid 90:10 and 95:5 partially overlap with each other. appear at around 10% carbon content. Rather, our results
scattered around the dotted line predicted by the ROM.
loaded to failure under tension to determine their It was indicated by some earlier studies [1,16,18] that
residual strength. For convenience of identification, the largest hybrid effects have been observed for hybrid
hereafter in the paper, specimen tested in air is referred composites with the highest degree of fiber dispersion.
to as ’dry’ specimen whereas specimen tested in distilled Bridging of a failed fiber bundle (Fig. 5) would be much
water is referred to as ‘wet’ specimen. more easier than that of an entire ply, thus preventing a
crack arising from the failed bundle from developing
into catastrophic one. Consequently, the tensile strength
3. Results and discussion of intraply hybrids presents a strong positive deviation
from the constant strain model, which only provides a
3.1. Quasi-static test lower bound for the tensile strength of hybrid composite.

Typical stress–strain curves for various hybrid compo- 3.2. Fatigue behavior
sitions are shown in Fig. 3. Tensile test results are sum-
marized in Fig. 4. All specimens (all-carbon, all-glass, Results of fatigue test of all-glass 100:0 and hybrid
and hybrid) show linear elastic behavior to final failure. 75:25 in air are summarized in Fig. 6. When calculating
Different tensile failure modes are seen, they varied from the slope of stress life (S–N) curves, test results of all-
dominant fiber roving splitting and bundle failure in all- glass 100:0 at 30%, and 45% UTS, and hybrid 75:25 at
glass specimens to dominant brittle-like failure mode in 45% UTS were excluded since at these load levels, all
all-carbon specimens where only few splitting between specimens survived 107 cycles and showed no sign of
fiber roving was seen. The tensile failure modes for damage as reflected in their residual strength (Figs. 7
hybrid specimens were in between these two extremes,

Fig. 5. Bridging of a broken carbon fiber bundle as indicated by the


Fig. 4. Experimental results and comparison with the constant strain black arrow. The black colored regions are carbon-fiber bundles; the
model and rule-of-mixture. white colored region is glass bundles.
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 851

composites of all kinds are linear and of a slope approxi-


mately 10% of the fracture strength. The lower slope in
our case compared with the Mandell stipulation could
be attributed to better fiber alignment or better fiber in
general. The slope of hybrid 75:25 S–N curve is 23.1
MPa per decade, or 5.1% of the static strength per dec-
ade. Thus, our hybrid composite showed an enhanced
fatigue resistance compared to all-glass specimens.
As has been well documented, change in elastic
modulus is a strong indicator of the state of the material
[19]. Data collected from various stress levels and dis-
similar fiber compositions suggest that the generic pat-
tern of modulus degradation during fatigue of for fiber
reinforced composites is independent of cyclic load level
and fiber compositions. This similarity reflects a resem-
blance in the essence of damage development process.
Fig. 6. S–N curves of all-glass and hybrid 75:25 in air. Some typical degradation curves of tensile modulus are
shown in Fig. 9, where tensile modulus and applied
cycles are presented in a normalized scale. A typical
curve is characterized by a small drop within the first
10% of life, followed by a plateau region extending to
about 80% of life, then succeeded by a much faster drop

Fig. 7. Residual strength of all-glass specimens after 107 cycles.

and 8). The S–N curve for the all-glass 100:0 specimens
has a slope of 18.8 MPa per decade of life, or about
5.7% of the static strength per decade. Our results do
not agree with the arguments of Mandell et al. [7,8],
who maintained that the slope of S–N curves for GRP

Fig. 9. Typical normalized stiffness change during fatigue in air: (a)


all-glass specimens; (b) hybrid 75:25 specimens. Arrow denotes run-
Fig. 8. Residual strength of hybrid 75:25 specimens after 107 cycles. out.
852 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

during the last stage of life, a damage related pattern that


is typical of fiber-reinforced composites.
At lower stress levels, some specimens showed a
slight increase in tensile modulus during fatigue loading.
Similar phenomenon was also reported elsewhere for
pultruded glass/vinyl ester composite [15]. A plausible
explanation is that the applied stress is unable to produce
enough fiber breakage and matrix cracking, which are
the main causes for the drop in modulus in the initial
stage. Instead, realignment of fibers and macromolecular
chains of the polymeric matrix along the direction of
applied stress may occur during cyclic loading, contribu-
ting to the increased modulus. However, it should be
noted that similar modulus enhancing mechanism may
also happen in high cyclic stress levels but it is over-
whelmed by the deleterious effect of fiber breakage and Fig. 10. Fatigue and environmental fatigue S–N data of all-glass
matrix cracking, resulting in an overall drop in the (100:0) specimens.
modulus. Residual strength test results of specimen sur-
vived 107 cycles are given in Figs. 7 and 8. Only hybrid
75:25 specimens cyclically loaded at 65% UTS showed
a 9% reduction in mean tensile strength while most data
did not show a significant reduction in strength, and
some specimens even showed a slightly increase which
may also attribute to the fiber realignment during the
fatigue process.
Macroscopic fatigue failure modes of glass 100:0 and
hybrid 75:25 were also examined. Failed specimens of
both types showed extensive longitudinal splitting within
the glass fiber bundles, possibly due to weaker interfacial
bonding between glass fiber and epoxy matrix. When
glass–carbon hybrid specimens were subjected to
fatigue, fiber breakage usually occurs initially within the
carbon bundles since at the same applied strain level
much higher stress is sustained by the carbon fiber bun-
dle than the glass fiber bundle. Subsequent to failure of Fig. 11. Fatigue and environmental fatigue S–N data of hybrid 75:25
carbon fiber bundle, catastrophic failure of the composite specimens.
may occur. For some hybrid specimens cyclically loaded
at lower stress levels, remaining intact glass and carbon
fiber bundles were still able to sustain the applied load dominant, which has also been shown previously in
until final failure after failure of the first carbon fiber another study for E-glass/vinyl ester [15]. A significant
bundle. This is reflected in the modulus degradation difference in fatigue life begins to emerge for those
curve as a stepwise drop toward the end of its fatigue tested at lower cyclic load levels. At 65% UTS, most
life [Fig. 9(b)]. dry specimens have failed around 106 cycles whereas the
fatigue lives of most wet specimens were clustered
3.3. Environmental fatigue around 105 cycles. The discrepancy in fatigue lives
between dry and wet specimens further widened at 45%
The environmental fatigue S–N data of all-glass and UTS. At that stress level, all dry specimens survived 107
hybrid specimens are presented in Figs. 10 and 11, cycles and showed no significant sign of damage because
respectively. For all-glass specimens cyclically loaded at there was essentially no change in residual tensile
85% UTS, the S–N data are indistinguishable between strength compared to as-fabricated specimens (Fig. 7).
the dry and wet specimens (Fig. 10). Due to the scat- However, a remarkable reduction in fatigue life is seen
tering of the experiment data and the relative short time for wet specimens, as their fatigue lives only ranged
of cyclic loading ranging from a few seconds to a few from 105 to 106 cycles, at least a decade shorter than
minutes, influence of environmental effect on fatigue those of dry samples. At 30% UTS all dry specimens
life, if there is any, may not be detectable. At higher and most of the wet specimens survived 107 cycles,
stress levels the environmental fatigue behavior is stress whereas a considerable difference in residual strength
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 853

was observed (Fig. 12). The average residual strength of


dry specimens showed virtually no change compared to
control, however, average residual strength of wet speci-
mens showed about 10% reduction. Therefore it is clear
that environmental fatigue behavior for all-glass speci-
mens is stress corrosion dominant at lower stress levels.
For the hybrid specimens, the situation is less drastic
in terms of difference in fatigue lives between wet and
dry specimens (Fig. 11). Similar to glass fiber speci-
mens, cyclic loading at 85% UTS did not produce sig-
nificant difference in fatigue lives between wet and dry
specimens. At 65% UTS, despite some overlapping of
data for dry and wet specimens, some of the dry speci-
mens survived to longer lives. Separation of fatigue life
for wet specimens from dry specimens become evident
Fig. 13. Stiffness reduction at 65% UTS.
only at 45% UTS. All dry specimens survived 107
cycles, and showed no sign of damage as is reflected
from the results of residual strength test, shown in Fig. specimens, indicating improved resistant to environmen-
8. However, all wet specimens failed before reaching 107 tal fatigue. At 65% UTS, wet hybrid specimens showed
cycles. Thus environmental fatigue of hybrid composite a life between 1.5×105 and 2.6×106 cycles, and for dry
is similar to that of all-glass composite, with stress domi- specimens 2.3×105 and 107 cycles. Both dry all-glass and
nant behavior at higher stress levels and stress corrosion hybrid samples survived 107 cycles at 45% UTS. How-
dominant behavior at lower stress levels. ever, fatigue lives of wet glass composite specimens
The similarity of data trends presented in Figs. 10 and ranged from 1×105 to 1×106 cycles, compared to
11 is that both all-glass and hybrid samples showed fas- 4.5×105–6.2×106 cycles for that of wet hybrid speci-
ter fatigue degradation in water compared to those tested mens, a 500% difference. Therefore, it is shown that the
in air. The influence of an aqueous environment on the resistance to environmental fatigue is substantially
tensile modulus of all-glass and hybrid specimens during enhanced during long-term cyclic loading by using only
cyclic loading is shown in Fig. 13. Both all-glass and 25% carbon fiber in hybrid composites compared to all-
hybrid wet specimens showed faster modulus degra- glass composites. In other words, hybridization results
dation than dry specimens. Hybrid specimens showed in a reduction of the stress corrosion dominant effect at
delayed degradation in fatigue life. At 65% UTS, no lower stress levels.
overlapping in fatigue lives is seen for all-glass com- When a specimen is cyclically loaded in water, water
posite specimens. Fatigue lives ranged from 5×103 to ingress in the specimen may involve capillary diffusion
1.5×105 cycles for wet specimens, and from 3.5×105 to of moisture along fatigue cracks as well as moisture
2×106 cycles for dry specimens. However, partial over- sorption from the specimen surface. However, with the
lapping in fatigue lives has occurred in the case of hybrid limited testing time involved at higher load levels, Fick-
ian absorption behavior may not be significant at room
temperature [10]. Thus, fluid ingress by way of matrix
cracking and fiber debonding may be a dominant mech-
anism [15]. Microscopic examinations on the fatigue
specimen’s surface revealed matrix cracking networks
and fiber debonding (Fig. 14) and it is believed that they
were the main passage for fluid ingress. It has been
shown that by direct weighting the fatigue samples the
sorption rate was found to be much faster than normal
Fickian sorption rate under load free condition [15].
However, the present study failed to reveal a similar
trend. Loss of material from the end-tab areas during the
frequent gripping and ungripping and possible transport
of material out to the fluid environment caused much
variation and error in data.
Previous studies suggested that the fiber/matrix
interphase region has a controlling effect in the environ-
Fig. 12. Difference in residual strength of all-glass specimens sub- mental fatigue performance of FRPs [20–24]. If the
jected to 107 cycles under 30% UTS. fiber/matrix interphase region is damaged or destroyed
854 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

Fig. 14. Transverse matrix cracking and fiber debonding on specimen Fig. 16. Fatigue failure surface of a specimen tested in water.
surface.

it is the reaction between water and the network mod-


ifying oxides of the glass that causes the reduction in
by fluid ingress, fiber/matrix debonding may propagate tensile strength [25]. Because carbon fibers are resistant
more easily, and fibers may lose their reinforcing func- to water and dilute acid, glass–carbon hybrid composites
tion. Examinations of the fracture surface under SEM as a whole are more resistant to stress corrosion than all-
revealed a difference on the fiber surface. Typically, glass fiber composites. By hybridization with high
more matrix residue adhering to the fibers is seen from modulus carbon fibers, the modulus of the hybrid is
dry specimens (Fig. 15) than that of wet specimens (Fig. increased while the overall failure strain is reduced.
16). This difference on the fracture surface indicates that When subjected to a same stress level, the stress carried
fluid action degrades the adhesion between the fibers and by the glass fibers is reduced. For instance, when loaded
the matrix. Thus the process of environmental fatigue at 65% UTS, all-glass specimen has a maximum nominal
involves coupled interaction between interphase degra- strain of 0.98% (212 MPa), and the tensile stress
dation and matrix cracking: once cracks in the matrix developed in the glass fiber is about 710 MPa, calculated
are developed, ingress of corrosive fluid into the material from the rule of mixture. However, at the same stress
is at a much faster rate, which further accelerates the level, applied tensile strain for hybrid 75:25 specimen is
degradation process of the interphase region. This reduced to 0.88% (296 MPa) and the stress shared by
coupled process exposes the glass fibers to corrosion at glass fiber is 640 MPa. Thus, the stress in the glass fiber
an increasing rate. is reduced by as much as 70 MPa or 11% for the hybrid
Early studies on the stress corrosion and sub-critical 75:25 samples. Since the subcritical crack growth rate
crack growth of soda-lime glass in water suggested that are shown to be a power function [26–29] of the applied
stress, it will notably reduce the corrosion rate of glass
fibers, and hence prolonging the fatigue life of the com-
posite. Furthermore, reduction in the composite failure
strain may also delay the onset of matrix cracking, there-
fore preventing the formation of passage for moisture
induct and postpone the fluid ingress process. As a
consequence, degradation of environmental fatigue life
is delayed substantially.

3.4. A simple life-prediction model

Being able to predict the fatigue response of a hybrid


composite from its parent materials is crucial in design-
ing such materials for optimized performance. In this
section, we present a simple phenomenological life pre-
diction model for hybrid unidirectional composites.
The hybrid composite is assumed to be composed of
Fig. 15. Fatigue failure surface of a specimen tested in air. two components with dissimilar properties, and the
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 855

fatigue behavior of each component is known empiri- N2


cally. The aim of this model is to utilize known fatigue D⫽ (9)
N1
data of parent materials to predict the fatigue behavior
of their hybrids with different volume ratio combi- Otherwise, if N1⬍N2, component 1 fails first, and the
nations. Two assumptions are made for the model: damage build up in component 2 is
N1
1. Each component works separately in the hybrid com- D⫽ (10)
N2
posite, and their behavior within the hybrid follows
the fatigue S–N curves of its pure status; In the second stage, after failure of one of its compo-
2. A linear damage theory of fatigue is applicable to nent, the remaining component may still able to sustain
each component, the applied load, sr, until final failure.

冘i
ni
Ni
ⱖ1 (3) sr ⫽
s·Er
Ernfr
(r ⫽ 1 or 2) (11)

where ni is the number of cycles experienced at a certain And the number of cycles to failure under sr is
stress level, si, and Ni is the number of cycles to failure Nr ⫽ f⫺1
r (sr) (r ⫽ 1 or 2) (12)
at si.
Taking into account the damage accumulated during
The S–N curve of a composite is most simply and stage 1, the remaining cycles to failure, Nremain, is
conveniently described by the following simple linear Nremain ⫽ Nr(1⫺D) ⫽ f⫺1
r (sr)(1⫺D) (13)
function
By adding the cycles in stage 1 and stage 2 together,
s⫽sult⫺msultlogNf (4) the total number of cycles to failure, Ntotal, is
where sult stands for its ultimate strength, m is the Ntotal ⫽ min(N1,N2) ⫹ Nremain (14)
decrease in fatigue strength per decade of fatigue life,
and Nf is the number of cycles to failure at applied cyclic where min (N1,N2) denotes the smaller value among N1
stress, s. However, since Eq. (4) may not be applicable and N2. Therefore, Ntotal is a function of applied stress,
to certain types of composites such as aramid fiber s, and the volume fraction, nf, of the component.
reinforced composites, a more general form would be Similar to the constant strain model in the static ten-
sile test of hybrid composites, this simple fatigue model
s ⫽ f(Nf) (5) does not take into account the synergetic effects such as
and in reverse form bridging of cracks in the low elongation component by
high elongation counterpart, thus deviations of actual
Nf ⫽ f⫺1(s) (6) fatigue response from the model predictions can be
regarded as fatigue hybrid effects. Thus if the actual
The loading history of a hybrid composite is divided fatigue life is longer than the prediction, hybridization
into two stages. The first stage is the period from the has a desirable positive synergetic effect. If the opposite
start of the test to the moment when one of its load- is true, then hybridization only yields a detrimental syn-
carrying components fails. Stage two is the duration ergetic effect or negative effect. In the following, model
from the failure of one of the components to final failure prediction will be compared to fatigue data for two types
of the composite. of hybrids, namely, the interply and the intraply com-
In the first stage, before a component fails, both posite.
components are able to carry the applied load, thus the
stresses s1, and s2, carried by components 1 and 2 are
3.4.1. Interply hybrids
s·E1 s·E2 Both all-carbon and all-glass unidirectional com-
s1 ⫽ s2 ⫽ (7)
E1nf1 ⫹ E2nf2 E1nf1 ⫹ E2nf2 posites essentially follow a linear S–N curve [Eq. (4)].
Fatigue behavior of unidirectional laminated carbon and
where s is the applied stress, E is the Young’s modulus, glass composites (in which carbon-fiber and glass-fiber
nf is the fiber volume ratio, subscript 1 or 2 denotes laminae are arranged in an alternate manner) are
component 1 or 2, respectively. Thus the number of extracted from [6], and are tabulated in Table 1. Predic-
cycles to failure for each component are tion of the fatigue response of a family of glass–carbon
N1 ⫽ f⫺1 laminated hybrids is illustrated in Fig. 17 by a 3-dimen-
1 (s1) and N2 ⫽ f2 (s2)
⫺1
(8)
sional (3-D) plot. Cycles to failure, and carbon fiber con-
If N1⬎N2 then component 2 fails first. Damage, D, tent are represented by the Y- and X-axis, respectively,
accumulated in the surviving component 1 is and applied cyclic stress is expressed by the Z-axis. As
856 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

Table 1
S–N curve parameters for all-glass and all-carbon unidirectional lami-
nated composites. From [6]

sult (GPa) m E (GPa)

UD carbon composite 1.97 5.7% 137.5


UD glass composite 1.30 9.8% 41.7

Fig. 18. Normalized fatigue curves of glass–carbon laminated


interply hybrid composites.

ally the first one to fail because it has a lower failure


strain compared to the glass-fiber component. Thus,
although glass component has an inferior fatigue
Fig. 17. Model predictions of fatigue curves of glass–carbon lami- response, it is being cycled at a much lower applied
nated interply hybrid composites. strain that causes relatively little damage and eventually
survives longer life. After failure of the carbon-fiber
component, the glass-fiber component usually suffered
illustrated in Fig. 17, the interception of the surface instantaneous failure. Only at two extreme conditions
graph to Plane 1 is the hybrid tensile strength predicted could it sustain more applied cycles, that is, when the
by the constant strain model. The lowest strength is
material is under very low fatigue stress or it has low
located around 30% carbon fiber content, and produces
carbon fiber content. Under these conditions, the glass
a ‘valley’ in the 3-D plot. To the right side of the valley,
component becomes dominant, which is indicated in Fig.
the fatigue behavior is dominated by carbon-fiber
18 as the steep transition region occurs from 0 to 30%
component, and to the left side, the glass-fiber compo-
carbon fiber content. Thus the fatigue response of glass–
nent. The S–N curve is also linear for various carbon
fiber contents to the right side of the valley with steeper carbon hybrid composites does not follow ROM, rather
slope at higher carbon fiber content. This may explain it seems to be carbon-component dominated for the most
the data reported by Dickson [6], as the author noted part. By hybridization with only a small fraction of car-
that the slope of the hybrid S–N curves are all roughly bon, the fatigue properties of the hybrid are altered to be
the same and as ‘unexpectedly’ lower than that of the similar to that of the all-carbon component. Therefore,
plain CFRP laminates. Also, it can be seen from the hybridization is a viable and cost-effective way to
graph that the fatigue stress varies gradually from the improve the fatigue performance of GFRPs.
constant strain model at low cycles to almost following Comparisons between the model prediction and
ROM at high cycles. experimental results of interply hybrid composites from
When normalized by their tensile strength, the fatigue [6] are shown in Fig. 19. Given the scatting nature of
behavior of glass–carbon hybrids becomes conspicu- fatigue behavior, the model is able to provide predictions
ous—it is carbon fiber dominated. As shown in Fig. 18, on the trend of the fatigue response of glass–carbon
normalized fatigue behavior of hybrids with carbon fiber laminated hybrid composites, although at high cycles the
content ranging from 30 to 100% show the same experimental data showed positive deviations from pre-
response, indicated by the flat surface on the graph. Cal- diction. These positive deviations may be attributed to
culations from the model revealed that in glass–carbon the crack arrest mechanisms by long elongation glass
hybrid composites, the carbon-fiber component is usu- plies.
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 857

value given by the constant strain model, thus the entire


experimental S–N data of hybrid 75:25 is higher than
the model prediction, which is an extrapolation of the
constant strain model. Therefore the proposed model can
be used as a lower bound for intraply fatigue behavior.
The fatigue behavior of both the intraply and interply
hybrid composites are examined. The difference between
the two is that the intraply hybrids showed a greater
deviation from model prediction, indicating a more
prominent hybrid effect during fatigue. More refined
models applying micromechanics and taking into
account the crack arresting mechanism are desirable. As
Fig. 19. Comparisons between model predictions and experimental mentioned earlier, carbon fiber usually fails first during
data for glass–carbon laminated hybrid composite. All experimental fatigue, the breakage may cause subsequent premature
data are extracted from [6]. failure in its vicinity. Bridging of a carbon fiber bundle
is much ‘easier’ than that of an entire carbon ply and
causes relative little damage to the adjacent areas. Thus
Table 2
S–N curve parameters for all-glass and all-carbon intraply unidirec- the damage development is slower than the interply
tional composites hybrid composites, and consequently it has a better
fatigue resistance.
sult (MPa) m E (MPa)
3.4.3. Predictions on environmental fatigue
UD carbon composite 710.5 4.7% 67.8 If the environmental fatigue behavior of the glass-fiber
UD glass composite 326.9 5.7% 23.4
component in the hybrid composites is assumed to be
the same as that of the environmental fatigue S–N curve
established for the all-glass composites, and the same is
3.4.2. Intraply hybrids also applicable to the carbon-fiber component, then the
Parameters used in the prediction of intraply hybrid environmental fatigue behavior of the glass–carbon
composites were extracted from our experimental data hybrid can be estimated by our fatigue model. Since it
described previously. The value m of carbon fiber poly- has been shown by previous studies that moisture does
meric matrix composites usually ranges from 0.035 to not alter the fatigue properties of CFRPs [11], we
0.059, thus an average value of 0.047 was adopted in assume that both fatigue and the environmental fatigue
our estimation based on data from the available literature S–N curve are the same for the all-carbon fiber
[9]. The strength, modulus, and m value of both all-car- reinforced composite. Based on these assumptions, the
bon and all-glass composites are summarized in Table 2. model prediction for environmental fatigue of an intraply
Comparison of the prediction of hybrid 75:25 and actual hybrid composite is shown in Fig. 21. As revealed in
experimental results are presented in Fig. 20. The experi- the figure, the environmental fatigue curve of hybrid
ment data show a considerable positive deviation from 75:25 can be divided into two regions. At low cycles,
the prediction. As discussed in the previous section the the curve is exactly the same as that of the fatigue curve
tensile strength of the intraply hybrids is higher than the for dry specimens, that is after the failure of the carbon-
fiber component, the hybrid composite suffers instan-

Fig. 20. Comparison between fatigue model prediction and the Fig. 21. Comparison between model predictions of environmental
experimental results of intraply hybrid composite. fatigue behavior of hybrid 75:25 and experimental results.
858 Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859

taneous failure. However, at high cycles, it is the glass- [2] Chou TW, Kelly A. Mechanical properties of composites. Annual
fiber component that fails first, followed by carbon fib- Review of Materials Science 1980;10:229–59.
[3] Phillips LN. The hybrid effect—does it exist? Composites
ers. Thus the fatigue response of a hybrid is dominated 1976;7:7–8.
by glass fibers at low stress for high cycle environmental [4] Marom G, Harel H, Neumann S, Friedrich K, Schulte K, Wagner
fatigue. This may further explain why the moisture HD. Fatigue behavior and rate dependent properties of aramid
effects become obvious only at low stress levels. The fibre/carbon fibre hybrid composite. Composites
experiment data show a considerable positive deviation 1989;20(6):537–44.
[5] Fernando G, Dickson RF, Adam T, Reiter H, Harris B. Fatigue
from the prediction. As discussed earlier, the tensile behavior of hybrid composites: part 1 carbon/kevlar hybrids.
strength of the intraply hybrids is higher than predictions Journal of Materials Science 1988;23:3732–43.
from the constant strain model, thus the entire experi- [6] Dickson RF, Fernando G, Adam T, Reiter H, Harris B. Fatigue
mental S–N data of hybrid 75:25 is higher than the con- behavior of hybrid composites: part 2 carbon–glass hybrids. Jour-
stant strain based life prediction model. This is also true nal of Materials Science 1989;24:227–33.
[7] Mandell JF, Huang DD, McGarry FJ. Tensile fatigue perform-
under fluid immersion condition. Therefore the proposed ance of glass fiber dominated composites. Composites Tech-
model can also be used as a lower bound for intraply nology Review 1981;3(3):96–102.
environmental fatigue behavior. [8] Mandell JF, McGarry FJ, Hsieh AJY, Li CG. Tensile fatigue of
glass fibers and composites with conventional surface compressed
fibers. Polymer Composites 1985;6(3):168–74.
[9] Liao K, Schultheisz CR, Hunston DL, Brinson LC. Long-term
4. Conclusions durability of fiber-reinforced polymer-matrix composite materials
for infrastructure applications: a review. SAMPE, Journal of
Advanced Materials 1998;30(4):2–40.
Base on the results obtained in this study, the follow- [10] Vauthier E, Abry JC, Bailliez T, Chateauminois A. Interactions
ing major conclusions can be reached: between hygrothermal ageing and fatigue damage in unidirec-
tional glass/epoxy composites. Composites Science and Tech-
앫 Cyclic loading in distilled water significantly shortens nology 1998;58:687–92.
[11] Jones CJ, Dickson RF, Adam T, Reiter H, Harris B. The environ-
the fatigue life of both unidirectional glass and glass– mental fatigue behaviour of reinforced plastics. Proceedings of
carbon hybrid fiber reinforced epoxy matrix com- the Royal Society London Series A: Mathematical and Physical
posites compared to their fatigue lives in air, at lower Sciences 1984;396:315–38.
stress levels. [12] Sumsion HT, Williams DP. Effect of environment on the fatigue
앫 Glass–carbon hybrid specimens showed a better of graphite–epoxy composites. Fatigue of composite materials,
ASTM STP 569: American Society for Testing and Materials,
retention in fatigue life in water than that of all-glass 1975, pp. 226–47.
composite specimens, up to 107 cycles. Therefore by [13] Komai K, Minoshima K, Shiroshita S. Hygrothermal degradation
hybridization with an appropriate amount of carbon and fracture process of advanced fibre-reinforced plastics.
fibers, resistance to environmental fatigue degradation Materials Science and Engineering A 1991;143:155–66.
of GFRP can be enhanced significantly. [14] McBagonluri F, Garcia K, Hayers M, Verghese KNE, Lesko JJ.
Charaterization of fatigue and combined environment on dura-
앫 The proposed phenomenological fatigue model is cap- bility performance of glass/vinyl ester composite for infrastruc-
able of estimating the fatigue behavior of hybrid com- ture applications. International Journal of Fatigue 2000;22:53–64.
posites from their parent composites. The fact that this [15] Liao K, Schultheisz CR, Hunston DL. Long-term environmental
simple model can better describe the behavior of fatigue of pultruded glass-fiber-reinforced composites under
interply composite suggests that synergistic effects in flexural loading. International Journal of Fatigue 1999;21:485–
95.
the intraply composites are significant and a refined [16] Hofer KE Jr, Stander M, Bennett LC. Degradation and enhance-
model entails the incorporation of the bridging mech- ment of the fatigue behavior of glass/graphite/epoxy hybrid com-
anism. posites after accelerated aging. In: Proceedings of the 32nd
Annual Technical Conference, The Society of the Plastics Indus-
tries, Inc. 1977, section 11-F.
[17] Manders PW, Bader MG. The strength of hybrid glass/carbon
Acknowledgements fiber composites, part 2. Journal of Materials Science
1981;16:2246–56.
The authors would like to acknowledge the support [18] Peijis AAJM, De Kok JMM. Hybrid composites based on poly-
ethylene and carbon fibres: part 6: tensile and fatigue behavior.
of this research by Nanyang Technological University Composites 1993;24(1):19–32.
through AcRF grant RG7/98. [19] Reifsnider KL, Schulte K, Duke Jr JC. Long-term fatigue
behavior of composite materials. In: O’Brien TK, editor. Long-
term behavior of composites, ASTM STP 813. Philadelphia:
American Society for Testing and Materials; 1983, p. 136–59.
References [20] Hofer KE, Skaper GN, Bennett Effe LC. Effect of moisture on
fatigue and residue strength losses for various composites. Jour-
[1] Manders PW, Bader MG. The strength of hybrid glass/carbon nal of Reinforced Plastics and Composites 1987;6:53–65.
fibre composites: part 1: failure strain enhancement and failure [21] Watanabe M. Effect of water environment on fatigue behavior
mode. Journal of Materials Science 1981;16:2233–45. of fiberglass reinforced plastics. In: Tsai SW, editor. Composite
Y. Shan, K. Liao / International Journal of Fatigue 24 (2002) 847–859 859

materials: testing and design (5th Conference), ASTM STP 674. on the Fracture Mechanics of Ceramics, vol. 4, Crack Growth
Philadelphia: American Society for Testing and Materials; 1979, and Microstructure. New York: Plenum Press; 1973, p. 549–81.
p. 345–67. [26] Evans AG, Wiederhorn SM. Proof testing of ceramic materials—
[22] Fried N. Degradation of composite materials: the effect of water an analytical basis for failure prediction. International Journal of
on glass-reinforced plastics. Mechanics of composite materials. Fracture 1974;10:379–92.
In: Wendt FW, Liebowitz H, Perone N, editors. Proceedings of [27] Wiederhorn SM. Subcritical crack growth in ceramics. In: Bradt
the Fifth Symposium on Naval Structure Mechanics, 1967, pp. RC, Hasselman DPH, Lange FF, editors. Symposium on the Frac-
813–37. ture Mechanics of Ceramics, vol. 2, Microstructure, Materials and
[23] Hofer KE Jr, Bonnett LC, Stander M. Effect of various fiber sur- Applications. New York: Plenum Press; 1973, p. 613–46.
face treatments on fatigue behavior of glass fabric composite in [28] Ritter Jr JE. Mechanics of subcritical crack growth in glass. In:
high humidity environments. In: Proceedings of the 31st Annual Bradt RC, Hasselman DPH, Lange FF, editors. Symposium on
Technical Conference of SPI, Washington DC, February 1976, the Fracture Mechanics of Ceramics, vol. 4, Crack Growth and
section 6A. Microstructure. New York: Plenum Press; 1973, p. 667–86.
[24] Shih GC, Ebert LJ. The effect of the fiber/matrix interface on the [29] Gupta PK. Examination of the tensile strength of E-glass fiber in
flexural fatigue performance of unidirectional fiberglass com- the context of slow crack growth. In: Bradt RC, Hasselman DPH,
posites. Composites Science and Technology 1987;28:147–61. Lange FF, editors. Symposium on the Fracture Mechanics of Cer-
[25] Wiederhorn SM. Mechanics of subcritical crack growth in glass. amics, vol. 5, Surface Flaws, Statistics and Microstructures. New
In: Bradt RC, Hasselman DPH, Lange FF, editors. Symposium York: Plenum Press; 1973, p. 291–303.

You might also like