You are on page 1of 14

Fatigue of Composite Materials

Ramesh Talreja
Department of Aerospace Engineering
Texas A&M University
College Station, Texas 77843-3141

Abstract This chapter summarizes part of the six lectures, pertaining to fatigue of
composite materials, presented at the session, "Modem Trends in Composite Lami-
nates Mechanics" at CISM in Udine. The summary provided here is of introductory
nature aimed at a reader who is not an expert in the subject. Ample references are
given to help the reader pursue the subject further.

1 Introduction
The subject of damage and failure of composite materials under static and time varying
loads is broad with many aspects ranging from the mechanisms involved to the mechanics
methods of analysis to failure criteria and life prediction. The six lectures prepared to
provide introduction and overview of the subject were organized as follows.
Lecture -1 Damage Mechanisms and Fatigue Life Diagrams- Part I
- Unidirectional Composites- PMCs
Lecture -2 Damage Mechanisms and Fatigue Life Diagrams- Part II
- Unidirectional Composites- MMCs and CMCs
- Fiber Architecture Effects
Lecture -3 Damage Mechanics
- Micromechanics vs. Continuum Damage Mechanics
- Damage Characterization
- Stiffness-Damage Relationships
Lecture -4 Damage Evolution
- Matrix Microcracking in Static and Fatigue Loadings
- Strength vs. Fracture Mechanics Approach
- A Mechanisms Based Model
Lecture -5 Damage and Failure from Notches
- Notch Stresses in Anisotropic Composites
- Failure Criteria
- Fatigue Damage From Notches
Lecture -6 Fatigue Testing and Evaluation
Cyclic Tension, Compression and Shear Testing
- Observation and Monitoring of Damage
H. Altenbach et al. (eds.), Modern Trends in Composite Laminates Mechanics
© Springer-Verlag Wien 2003
282 R. Talreja

Figure 1. A composite ply element subjected to in-plane stresses.

- Nondestructive Evaluation Techniques


This compendium will only cover the introductory aspects of fatigue damage mechanisms
and modeling. A full coverage of all topics listed here will appear in a book in preparation
("Damage, Fatigue and Failure of Composites", by R. Talreja and E.K. Gamstedt) to be
published by Cambridge University Press.

2 Damage Mechanisms and Fatigue Life Diagrams


2.1 Unidirectional Polymer Matrix Composites
When a piece of metal is subject to repeated loads, the irreversible set of mechanisms
commonly described as plasticity cause fatigue crack initiation and growth leading to fail-
ure. In a polymer matrix composite (PMC) with elastic fibers there is no plasticity but
other irreversible mechanisms can cause fatigue failure. These mechanisms, described
more appropriately as damage, can take different forms depending on the constituent
properties, their geometrical configuration and the type of loading. To provide a system-
atic framework for assessing the damage mechanisms, Talreja (1981) proposed fatigue life
diagrams. In later works Talreja (1993), Talreja (2000), Gamstedt and Talreja (1999)
further treatment of these diagrams has been presented. While the reader is urged to
look into these references for more details, the following summary will serve as a brief
introduction.
When a composite structure is subjected to a general time-varying load, e.g. a cylinder
under cyclic bending and torsion or a panel under cyclic biaxial tension, the failure of the
structure is often determined by failure of fibers in the critical load bearing ply. That ply
could in general be stressed by in-plane stresses as illustrated in Figure 1. The fatigue
failure of the critical ply can be viewed as occurring due to the primary mechanism of
fiber failure modified by other mechanisms such as matrix cracking and fiber/matrix
debonding. To facilitate a systematic assessment of the damage mechanisms and their
consequence on fatigue life, Talreja (1981) proposed a representation called fatigue life
diagram. This diagram is not the usual stress-life (S-N) diagram, commonly used for
Fatigue of Composite Materials 283

Region I

Region III

LogN

Figure 2. Fatigue life diagram for a unidirectional composite subjected to load-controlled


cyclic tension along fibers. Vertical axis is the maximum strain attained in the first
application of tensile load. Horizontal axis is the logarithm of the number of cycles to
failure.

metal fatigue, but is a mechanisms based plot of characteristic regions in the fatigue
behavior of a given composite under specific loading conditions. Figure 2 shows the
fatigue life diagram of the baseline case of a unidirectional composite under cyclic tension.
The three regions shown in Figure 2 represent the governing mechanisms within the
specified ranges of the composite strain. As explained in detail elsewhere Talreja (1981),
Talreja (1993), Talreja (2000), Region I operates within the scatter band of the composite
failure strain, etc., and the fiber failure mechanism here is non-progressive; thus the
failure can occur in any number of cycles. Region II represents the main progressive
mechanism of fiber failure occurring largely between the planes of a fiber-bridged crack
(Figure 3), but also in the parts of the volume affected by this cracking and aided by
fiber/matrix debonding, Gamstedt and Talreja (1999). Finally, Region III is the region
of damage that does not lead to failure in the preselected large number of cycles, e.g.
106 , and lies below the so-defined fatigue limit. The level of strain marked Em in the
diagram stands for the fatigue limit of the matrix material, as this limit is often also the
fatigue limit of the composite.
284 R. Talreja

Jl JUl
t~
Figure 3. Fiber-bridged matrix crack.

2.2 Unidirectional Metal Matrix Composites


The fatigue life diagram was conceived on the basis of observed and conjectured
damage mechanisms in polymer matrix composites. Later, some data on metal matrix
composites (MMCs) were analyzed and changes in the diagram relating to this material
system were proposed in Talreja (1995). More generally, certain trends in the diagram
were identified that were supported by the MMC data. Figure 4 shows the trends induced
by constituent properties. Note that increasing ductility decreases fatigue life in MMCs.
This is contrary to metals where ductility (lower yield stress) induces crack tip blunting,
thereby enhancing crack growth resistance. In MMCs crack tip blunting increases crack
opening displacement, which places increased tensile strain on the bridging fibers between
the matrix cracks, thereby failing fibers earlier. The effect of increasing fiber stiffness is,
on the other hand, beneficial, as indicated in Figure 4. Stiffer fibers tend to decrease
crack opening displacement by providing greater resistance to stretching displacement.
This, in effect, applies closing traction to the crack planes, reducing thereby the crack
growth rate. The trend in the fatigue limit follows the trend in the fatigue life.
All data analyzed for PMCs was found to support the nature of Region I in the
fatigue life diagram. Either direct support of the horizontal scatter band was found or
missing data in this range of strain could be inferred to lie in a horizontal scatter band.
However, the main condition for the non-progressive nature of fiber failure was argued to
be the insufficient availability of irreversible deformation in the highly constrained matrix,
see Talreja (1981). In an MMC the matrix will possess more irreversible deformation
(plasticity) than in a PMC, given equal conditions of fiber properties and volume fraction.
At high temperatures the yield stress of a metal decreases and strain softening occurs,
which would provide incentive for further deviation from the non-progressive fiber failure.
Thus Region I of the fatigue life diagram would slope away from the horizontal as the
temperature is increased. This was indeed supported by the data analyzed in Talreja
(1995).

2.3 Unidirectional Ceramic Matrix Composites


Typical ceramic matrix composites (CMCs) have ceramic fibers embedded in a ce-
ramic matrix. The bond between the two constituents is usually weak and one relies upon
their acting together by remaining in frictional contact after fiber/matrix debonding. Of-
Fatigue of Composite Materials 285

Region I

Fiber stiffuess

Fatigue
limit

Region III

LogN
Figure 4. Trends in the fatigue life diagram induced by constituent properties.

ten this aspect is considered in selecting the constituents, i.e. assuring that the thermal
expansion mismatch produces a compressive normal stress on the fiber/matrix interface.
The frictional sliding at the interface then forms the main irreversible mechanism that
plays a key role in fatigue of CMCs, as discussed in the following.
Before considering fatigue it is useful to ask what happens when a tensile axial load
is applied to a CMC. Let us first think of a load that is too low to cause any cracking.
Assuming the matrix failure strain is lower than the fiber failure strain, as is commonly
the case, the fibers remain intact at such a load level. If the same load is reapplied
after unloading, then the fibers would remain intact, as they would be strained to the
same value as in the first load application, assuming fibers as well as matrix are elastic
materials. This consideration suggests that the minimum condition to satisfy for fatigue
to occur is that the matrix failure strain (which is also the composite failure strain) is
reached. Once matrix cracking has occurred, several scenarios can be envisioned for fiber
failure progression after the first application of load. These scenarios are sketched in
Figure 5. The first scenario from left in Figure 5 corresponds to the case when the first
load has caused cracks that extend across the entire specimen cross section. The extent
of fiber failures in the first load will depend on the maximum strain level reached. Thus if
this strain is within the fiber failure scatter band, then the fiber failure will be extensive.
In this case the subsequent load repetition will not cause a progressive fiber failure, as a
small increment in fiber failure in any of the cross-sections with matrix cracks will lead to
286 R. Talreja

Figure 5. Three scenarios of matrix cracking and fiber failure in unidirectional CMCs.

total composite failure. This will result in the non-progressive Region I discusses above
for PMCs.
The second damage scenario inFigure 5 results when the first load level is high enough
to cause partial matrix cracks along with some fiber failures. When the load is repeated
the interfacial sliding in the debonded regions provides the irreversible mechanism re-
sponsible for stress redistribution in fibers. Thus, as the matrix cracks grow, more stress
redistribution is potentially available and progressively the fibers fail. It is likely that the
cracks become fully developed before total failure, if the load level is high enough. At
relatively low load levels, the matrix cracks may still be partially grown at the point of
unstable crack growth. The matrix cracking and the associated progressive fiber failure
form the damage of Region II.
The last scenario illustrated in Figure 5 is for the mechanism of Region III where
matrix cracks may form but their growth does not carry sufficient impetus for fibers to
fail at rates that could cause total failure in a large number of cycles. The fatigue limit
will depend on the specific value of that number; the larger the number, lower the fatigue
limit. An infinite number of cycles will correspond to the fatigue limit where no cracks
can be formed.

2.4 Fiber Architecture Effects


Composite structures are built by placing fibers in different orientations to effectively
carry multi-axial loading. We will here consider how the mechanisms of fatigue damage
are affected by multi-directional fiber placement in a laminate subjected to a cyclic tensile
load and how that translates into changing the basic fatigue life diagram described above.
We begin by considering the simple case of a single fiber orientation in a composite
placed at an angle r.t. to the applied cyclic tension load (Figure 6). As illustrated in
the figure, cracks initiate along the fibers, either at interfaces or in the matrix. At first
the cracks may initiate from defects in the matrix or at interfaces and grow with cycles
through the thickness as well as along the fibers. At some stage in the evolution of this
damage, a single crack may grow to the extent of attaining unstable growth in the next
application of load and separate the composite in two pieces, as illustrated in the figure.
In comparison to the on-axis loading case, i.e., when the fibers are aligned with
the loading direction, the damage mechanism of the off-axis loading case is drastically
Fatigue of Composite Materials 287

l
Figure 6. Cracking in a unidirectional composite under off-axis cyclic tension (left) and
failure from growth of a crack (right).

different and dramatically simple. The single progressive mechanism depicted in Figure
6 will produce a continuous curve (and the associated scatter band) in the fatigue life
diagram starting at the first-cycle failure strain and ending asymptotically in the fatigue
limit. For each off-axis angle the curve will be different. A schematic depiction of the
fatigue life diagram is shown in Figure 7, which illustrates the fatigue life dependence on
the off-axis angle. The on-axis fatigue life diagram is shown for reference in the figure in
broken lines. As Figure 7 illustrates, Region I (horizontal scatter band) of the fatigue
life diagram does not exist when the applied cyclic load is inclined to the fiber direction.
The single scatter band starts at the composite failure strain and asymptotically ends
at the fatigue limit. The failure strain as well as the fatigue limit strain depend on the
off-axis angle, e. A set of fatigue data for a glass/epoxy composite examined in Talreja
(1981) indicated that the fatigue limit strain varied from 0.6% at e = 0 to 0.12% at
e = 90. Most of the drop in the fatigue limit occurred until the off-axis angle of 30°.
Next we consider the case of two fiber orientations. In most practical cases, fibers are
placed symmetrically about the major loading axis, in which case the plies with fibers
in the two orientations will receive equal damage. The nature of this damage is different
from the case of a single fiber orientation. When a given ply cracks along its fibers, it
does not fail (separate in two pieces), as the adjacent ply it is bonded to holds it in place.
The adjacent ply cracks as well and is prevented from separation similarly. The stress
normal to the crack planes responsible for crack formation builds up over a distance called
shear lag (because the shear stress operates over this distance) and another crack can
form when the maximum normal stress reaches a critical value. This leads to multiple
cracking in the plies, as illustrated in Figure 8. When cracks in an off-axis ply are fully
developed, they span the ply thickness as well as the width of that ply along the fiber
288 R. Talreja

.. e
IIITIIIIITIT

""
~-~L.L L.L L.L L.L
(6=0)
-~
""'--~

""-~
(O<B<90)

LogN
Figure 7. Fatigue life diagram for unidirectional composites under off-axis loading. The
diagram in broken lines is for the on-axis loading case.

direction. As the cracks approach the ply interfaces, the intense stress field along the
crack fronts fails the interfaces. This ply separation, also called internal delamination,
initiates first in the regions where crack fronts of the two off-axis orientations intersect,
and progresses along the fronts of the two cracks. Figure 8 shows delamination along one
off-axis direction for clarity, leaving to visualize similar delamination in the other off-axis
direction. Failure occurs when the individual local strip-shaped delaminations widen and
coalesce with neighboring delaminations, causing large delaminated ply regions that fail
on overloading of fibers.
The fatigue life diagram for angle ply laminates have the same features as for the
off-axis loading case illustrated in Figure 7. However, the fatigue limit for angle ply
laminates shows considerable improvement with respect to the unidirectional off-axis
loading case, for off-axis loading angles lower than approximately 400, beyond which an
abrupt decrease of fatigue limit occurs, see Talreja (1981).
When the two fiber orientations are orthogonal, such as in a cross-ply laminate, and
loading is tensile along one of the fiber directions, the sequence of damage mechanisms
is as follows. First, cracks initiate in the transverse plies at locations that have defects
from manufacturing. These cracks usually span the thickness of the transverse plies
and then grow in the fiber direction, i.e., along the width of the transverse plies. This
Fatigue of Composite Materials 289

Delamination

Matrix
crack

Figure 8. Multiple matrix cracks in off-axis ply of a laminate (left) and subsequent
delamination caused by fatigue (right). Cracks and delamination are shown for one ply
only for clarity; the other ply is indicated in broken lines.

cracking process is non-interactive in the sense that the individual cracks do not influence
one another. When a few cracks have grown fully along the thickness and width of the
transverse plies, an interactive process sets in where neighboring cracks determine the
stress distribution in the ply region between the cracks, causing another crack to initiate
(midway) between them. As the applied load increases, the number of cracks per axial
length increases. When the maximum load in a cycle is repeated, the stress distribution
in the transverse plies changes only if an irreversible process exists. Akshantala and
Talreja (2000) identified this process to be frictional sliding between locally debonded
ply interfacial surfaces. Thus further transverse cracking occurs as number of load cycles
increases. Final failure occurs when the axial ply regions separated from the transverse
plies are overstressed.
The case of cross ply laminates has been studied extensively because it can be sim-
plified to a two-dimensional case for stress analysis. Various approximate analytical
solutions to the two-dimensional case can be found in Nairn and Hu (1994) where the
process of multiple cracking is also treated. The features of the fatigue life diagram are
similar to those of a unidirectional composite, illustrated in Figure 2. Region I exists for
cross ply laminates because at applied strains close to the failure strain of the composite
(which is also the failure strain of fibers), the axial plies are essentially the load bear-
ing plies, as the transverse plies at that strain would be far into their multiple cracking
process. The fatigue limit for cross ply laminates is determined by the strain to first
cracking of the transverse plies. Region II, lying between the fatigue limit and Region
I, represents the progressive mechanisms of multiple cracking in transverse plies and the
290 R. Talreja

associated local delaminations discussed above.


For practical structures the composite architecture is designed to satisfy multiple re-
quirements, such as resistance to bending and torsion as well as thermal expansion. A
common composite architecture is a laminate consisting of plies, each with unidirectional
fibers, stacked in a sequence such that the resulting structure has the required combina-
tion of properties. An example is [0/ ± 45/90] 8 laminate, which is quasi-isotropic, i.e.,
it has directionally independent average elastic moduli in its mid-plane. Many other
laminate configurations are possible, but often the number of ply orientations are kept
to three, and the most used ones are 0, 45 and 90. From fatigue point of view the assess-
ment of laminates is remarkably simple. First, irrespective of other ply orientations, the
presence of 0° plies provides Region I, which lies as a scatter-band about the fiber failure
strain, as in the unidirectional on-axis loading case as well as in the cross ply laminate
case. The fatigue limit for any laminate is determined by the first cracking mechanism.
Therefore, if the 90° ply orientation is present in a laminate, the strain at which first
transverse cracking occurs will determine the fatigue limit. This strain value is affected
by the so-called ply constraint, i.e., the ratio of the transverse ply modulus to the axial
modulus of the constraint-providing plies as well as the thickness ratio of the cracking
plies to the constraining plies. For further discussion of the constraint effect see Talreja
(1985). The progressive fatigue damage, represented by Region II, appears as a sloping
scatter band, starting at a low number of cycles (10 2 -103 ) and asymptotically approaching
the fatigue limit at a high number of cycles (10 6 -10 7 ). Region II displays relatively low
sensitivity to the orientation of the plies in a laminate. This aspect, although not fully
appreciated by designers, allows robust design against fatigue of composite laminates.
Note that the robustness feature is only evident when the composite fatigue behavior is
analyzed through the fatigue life diagram. The traditional S-N plotting of fatigue data
has prevented a proper appreciation of the composite fatigue behavior.

3 Modeling of Composite Fatigue


Modeling of the composite fatigue damage process serves several purposes. It aids in
developing materials with desired fatigue resistance, it helps in selecting existing material
systems to meet design requirements and it allows in making life assessment of structures
for safe performance. In general fatigue modeling must address three areas 1) The criteria
for initiation of fatigue damage, 2) The kinetics of fatigue damage evolution, and 3) The
criteria for criticality of fatigue damage, i.e., fatigue failure. We shall address each of
these areas below.

3.1 Fatigue Damage Initiation


Modeling of fatigue damage initiation is probably the most difficult aspect of fatigue
modeling. Even in monolithic materials, in particular metals, whose fatigue has been
a subject of study for more than a century, modeling of fatigue damage initiation still
remains a challenge. With the emergence of fracture mechanics the approach taken to
fatigue crack initiation has been to assume that a flaw (modeled as a crack) exists, and
initiation of fatigue has been cast in terms of a threshold that must be exceeded by a
Fatigue of Composite Materials 291

combination of load level and crack size for the existing crack to grow in fatigue. The
fatigue threshold for metals and other materials has been the subject of extensive studies
since the early 1g70s. In most cases an experimental determination of the threshold
is relied upon since a model-based prediction has not met with much success. In a
unidirectional composite loaded in cyclic tension along fibers, the threshold for fatigue
cracking in the matrix could be the starting point, beyond which consideration must
be given to how the presence of fibers in the path of a growing crack would affect the
threshold for composite fatigue. In doing this a fracture mechanics approach has not
been successful because of the complexities brought in by fiber bridging, fiber/matrix
debonding and the local irregularity of fiber distribution. The lack of self-similarity of
crack growth, required for the crack growth analysis by fracture mechanics, renders an
effort in this direction close to being futile. The alternative way is to simply extract
a fatigue threshold variable by experimental means. The appropriate variable, in the
context of the fatigue life diagram discussed in section 2 above, is the maximum first
cycle strain below which fatigue failure would not occur in a preselected large number of
cycles. Noting that the fatigue threshold strain for the matrix material must be exceeded
for a crack to initiate in the composite, this threshold value is a good estimate for the
fatigue limit strain for the composite. Obviously, the presence of broken fibers and
other flaws in the composite will affect this estimate. It turns out that for glass/epoxy
composite the fatigue limit strain of epoxy, which was found to be 0.6% by fatigue testing
of the unreinforced epoxy, holds fairly well for the composite as well at different fiber
volume fractions, see Dharan (1g75) and Talreja (1g81).
For laminates the threshold strain for fatigue initiation depends on the ply that would
crack first. For cross ply laminates and laminates which have goo plies, the strain to
initiation of transverse cracking provides the laminate fatigue limit. If goo plies are not
present, initiation of cracking in the most off-axis plies will govern the fatigue limit.

3.2 Evolution of Fatigue Damage


Let us begin with unidirectional composites first. Under cyclic tension parallel to
fibers, the progressive damage mechanism is the growth of fiber-bridged cracks, as de-
scribed above. This mechanism has been studied extensively for ceramic matrix com-
posites and to a lesser degree for metal matrix composites. Most studies have focused
on composite strength and fracture toughness aspects, while much less attention has
been given to crack growth in fatigue. Gamstedt and Talreja (1ggg) in a study of car-
bon/epoxy observed that fiber-bridged cracks emanated typically from broken fibers and
grew at decelerating rates. The extent of failure of bridging fibers was found to be small,
typically 5-8 fibers, and failure was found to occur from joining of fiber failed regions
through fiber/matrix interface failures. These conditions in polymer matrix composites
have frustrated efforts to model the fatigue damage evolution.
The situation is different in cross ply laminates where the progressive damage consists
mainly of transverse ply cracking. To properly model this damage one must explain the
multiplication of transverse cracks with cyclic loading. The fracture mechanics based
models dwell in correlating the crack density (number of cracks per unit axial length of
laminate) with applied load (via an energy release rate type of parameter), see Nairn
292 R. Talreja

and Hu (1994) and Hanaff-Gardin et al. (2000). This is no different from the fracture
mechanics approach to growth of a single crack where the rate of crack length increase
with cycles is correlated with a load parameter (range of stress intensity factor). An
alternative approach to the crack density increase in cross ply laminates under cyclic
loading was offered by Akshantala and Talreja (1998) where this increase was attributed
to an irreversible increase of stress in the transverse plies. The model put forward as-
sumed a frictional shear stress in the debonded 0/90 ply interface, which was responsible
for stress redistribution in the plies.
For general laminates the progressive fatigue damage process consists of multiplication
of cracking in all off-axis plies, generally at different rates depending on the constraints of
the cracking plies. The model development for this case is hindered by the unavailability
of analysis (even approximate) of the local ply stresses.

3.3 Final Fatigue Failure


Generally, any laminate is expected to fail when the final load-bearing plies fail. In
most laminates of multi-directional plies, one set of plies is aligned with the major load
direction. These plies are referred to as "axial" plies and the plies of other orientations
are called off-axis plies. The importance of the failure of axial plies is made evident in
Figure 9, where fatigue life data are plotted for [0/ ± 45] 8 and [0/ ± 45/90] 8 laminates
along with the fatigue life diagram for the unidirectional composite. Note that the fatigue
life data for both laminates fall within the 0° fatigue life diagram Region II at high cycles
and deviate from this region at low cycles. This suggests that at high cycles to failure,
corresponding to low applied strains, the off-axis plies transfer their share of the load to
the axial plies early in the fatigue process, resulting in the axial plies effectively providing
the laminate fatigue resistance. Most likely this occurs by the axial plies delaminating
from the off-axis plies early in the fatigue process. At high strains the fiber failures in
the axial plies may occur earlier because of the stress concentrations caused by matrix
cracking in the off-axis plies, while the interface between the axial plies and the off-axis
plies is still intact. Because of the importance of the axial ply failure we shall discuss
modeling of this process below.
Fiber failures in a unidirectional composite cause the final failure. The fiber failure
process is statistical, even for a single fiber. Classical models such as the Weibull strength
model, which assumes failure to come from the weakest link in the "chain" of fiber
elements, provide failure probabilities. When a family of parallel fibers is loaded, the
failure process is affected by the so-called "load sharing", which depends on the degree of
interaction between fibers in the family of fibers. If the fibers are not in contact and not
embedded in a matrix, the load sharing is equal, otherwise it is "local", i.e., determined
by the interactive conditions in the close vicinity of a broken fiber site. For the equal
load sharing case, the failure probabilities are given by a model by Daniels (1945), among
others. A range of models has been proposed for the local load sharing case, notably by
Rosen (1962) as the first such model, and then by several researchers. Current trend in
the model development is to make use of computer simulations of the failure progression.
Although failure progression in the monotonically increasing load case is important,
it is negligible in Region I of the fatigue life diagram. No models exist as yet to predict
Fatigue of Composite Materials 293

1.5 0

o Unidirectional
0.9 0 (0, ±45 )~
• ( 0, :t 45,90 )$

0.60~---~2-----+4-----!:6:----­
logN

Figure 9. Multiple matrix cracks in off-axis ply of a laminate (left) and subsequent
delamination caused by fatigue (right). Cracks and delamination are shown for one ply
only for clarity; the other ply is indicated in broken lines.

the occurrence of failure in this region. As argued above, the failure is determined by
the initial conditions, which govern the fiber failures in the first application of load.
Final failure from these initial failures occurs with "random chance" without significant
progression involved. In Region II the final failure comes from regions of fiber-bridged
cracks getting connected by interfacial debonds Gamstedt and Talreja (1999), or by
a single crack growing unstably. The role of interfaces was clarified in a comparative
study of different interfaces by Gamstedt et al. (1998). An account of the interfacial
characteristics in developing the final failure criteria thus appears to be important.

4 Conclusion
These notes have summarized the salient features of fatigue damage mechanisms in com-
posite materials, while providing a rational framework for interpreting fatigue life data
and utilizing it as guidance for modeling of the fatigue behavior. Beginning 1981, when
this author first introduced the interpretative framework for fatigue of composites, called
fatigue life diagrams, a tool has existed for assessing the roles of matrix and fiber mate-
rials, and of laminate configuration, in determining the composite fatigue performance.
However, the century old field of fatigue of metals has dominated much of the thinking of
researchers and engineers, and has slowed progress, as has isotropic materials experience
been a hindrance in composites development. These notes, written for newcomers to
294 R. Talreja

the field of fatigue of composites, are yet another effort to bring the much needed new
thinking to the this field.

Bibliography
Akshantala, N.V., and Talreja, R. (2000). A Micromechanics Based Model for Predicting
Fatigue Life of Composite Laminates. Materials Science and Engineering A285:303~
313.
Akshantala, N. V., and Talreja, R. (1998). A Mechanistic Model for Fatigue Damage
Evolution in Composite Laminates. Mechanics of Materials 29:123~140.
Dharan, C.K.H. (1975). Fatigue Failure Mechanisms in a Unidirectionally Reinforced
Composite Material. In Fatigue of Composite Materials, ASTM STP 569, Philadel-
phia: American Society for Testing and Materials, 171 ~ 188.
Daniels, H.E. (1945). Proc. R. Soc. Land. A183:405~435.
Gamstedt, E.K., Berglund, L.A., and Peijs, T. (1998). Influence of interfacial strength
on micro- and macroscopic behavior of longitudinal glass fiber reinforced polypropy-
lene. In K.L. Reifsnider, D.A. Dillard, and A.H. Cardon, eds., Progress in Durability
Analysis of Composite Systems. Rotterdam: A.A. Balkema, 137~ 142.
Gamstedt, E.K., and Talreja, R. (1999). Fatigue Damage Mechanisms in Unidirectional
Carbon-Fibre-reinforced Plastics. Journal of Materials Science 34:2535~2546.
Hanaff-Gardin, C., Goupillaud, 1., and Lafarie-Frenot, M.C. (2000). Evolution of matrix
cacking in cross-ply CFRP laminates: Differences between mechanical and thermal
loadings. In A.H. Cardon, H. Fukuda, K.L. Reifsnider, and G. Verchery, eds., De-
velopments in Durability Analysis of Composite Systems. Rotterdam: A.A. Balkema,
69~76.

Nairn, J.A., and Hu, S. (1994). Matrix Microcracking. In R. Talreja, ed., Damage Me-
chanics of Composite Materials. Amsterdam: Elsevier Science, 187~243.
Rosen, B.W. (1962). AIAA Journal2:1985-1991.
Talreja, R. (1981). Fatigue of Composite Materials: Damage Mechanisms and Fatigue
Life Diagrams. Proc. R. Soc. Land. A378:461~475.
Talreja, R. (1985). Transverse Cracking and Stiffness Reduction in Composite Laminates.
J. Camp. Mats 19:355~375.
Talreja, R. (1993). Fatigue of Fiber Composites. In T.W. Chou, ed., Materials Science
and Technology, Chapter 13, Weinheim: VCH, 584-607.
Talreja, R. (1995). A Cenceptual Framework for Interpretation of MMC Fatigue. Mate-
rials Science and Engineering A200:21~28.
Talreja, R. (2000). Fatigue of Polymer Matrix Composites. In R. Talreja and J.-A.E.
Manson, eds., A. Kelly and C. Zweben, Series eds., Comperehensive Composite Ma-
terials, Vol. 2, Oxford: Elsevier, 529~552.

You might also like